Hostname: page-component-55f67697df-zh294 Total loading time: 0 Render date: 2025-05-10T11:57:44.695Z Has data issue: false hasContentIssue false

A flower theorem in dimension two

Published online by Cambridge University Press:  08 May 2025

LORENA LÓPEZ-HERNANZ*
Affiliation:
Departamento de Física y Matemáticas, Universidad de Alcalá, Alcalá de Henares, Spain
RUDY ROSAS
Affiliation:
Departamento de Ciencias, Pontificia Universidad Católica del Perú, Lima, Peru (e-mail: [email protected])
Rights & Permissions [Opens in a new window]

Abstract

We prove a two-dimensional analog of the Leau–Fatou flower theorem for non-degenerate reduced biholomorphisms tangent to the identity.

Type
Original Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1 Introduction

Let $F\in \operatorname {\mathrm {Diff}}(\mathbb C^n,0)$ be a germ of a biholomorphism tangent to the identity. In dimension one, the dynamics of F is completely described by the Leau–Fatou flower theorem [Reference Fatou6, Reference Leau9], which guarantees the existence of simply connected domains with zero in their boundary, covering a punctured neighborhood of the origin, which are stable either for F or for $F^{-1}$ ; moreover, in each of these domains F is conjugated to the unit translation.

In dimension two, no complete description of the dynamics of F is known. Some partial analogs of the Leau–Fatou flower theorem have been obtained, guaranteeing the existence of either one-dimensional [Reference Abate1, Reference Écalle5, Reference Hakim8, Reference López-Hernanz and Rosas12] or two-dimensional stable manifolds [Reference Hakim7, Reference Vivas15]. With no extra assumptions on F, the most general result is due to Abate [Reference Abate1], who showed that F always supports some stable dynamics: either F has a curve of fixed points or there exist one-dimensional stable manifolds of F with the origin in their boundary.

The proof of the above-mentioned results is crucially based on a resolution theorem for F, which reduces the study of the dynamics of F to some combinatorial data of the resolution and the study of the local dynamics of some reduced models of the transform of F. This resolution theorem was introduced by Abate in [Reference Abate1] and is based on the corresponding result for vector fields due mainly to Seidenberg [Reference Camacho, Lins Neto and Sad4, Reference Seidenberg14]. Before stating the resolution theorem, we establish a precise definition of the reduced models suited to our purposes (see also Remark A.4).

Definition 1.1. The germ of a biholomorphism in dimension two at a fixed point p is called reduced if it is analytically conjugate to one of the following models.

  1. (i) Regular fixed points:

    $$ \begin{align*}\tilde{F}(x,y)=(x+x^{M}y^N[1+A(x,y)],y+ x^{M}y^{N}B(x,y)),\end{align*} $$
    where $M,N\ge 0$ , $(M,N)\notin \{ (0,0), (1,0)\}$ , $\operatorname {\mathrm {ord}} A\ge 1$ and $B\in (y)$ if $N\ge 1$ .
  2. (ii) Non-degenerate fixed points:

    $$ \begin{align*}\tilde{F}(x,y)=(x+x^{M+1}y^N[a+A(x,y)], y+x^My^{N}[by+B(x,y)]),\end{align*} $$
    where $M\ge 1$ , $N\ge 0$ , $ab\neq 0$ , ${a}/{b}\notin \mathbb {Q}_{>0}$ , $\operatorname {\mathrm {ord}} A\ge 1$ , $\operatorname {\mathrm {ord}} B\ge 2$ and $B\in (y)$ if $N\ge 1$ .
  3. (iii) Saddle-node fixed points:

    $$ \begin{align*}\tilde{F}(x,y)=(x+x^{M}y^N[x+A(x,y)], y+x^My^{N}B(x,y)),\end{align*} $$
    where ${M, N\ge 0}$ , $M+N\ge 1$ , $\operatorname {\mathrm {ord}} A, \operatorname {\mathrm {ord}} B\ge 2$ , $A\in (x)$ if $M\ge 1$ , $B\in (y)$ if $N\ge 1$ and $x+A$ and B have no common factors.

The resolution theorem, as we explain in Appendix A, guarantees the existence of a finite composition of blow-ups $\pi :(M,E)\to (\mathbb C^2,0)$ , with $E=\pi ^{-1}(0)$ , which transforms F into a map $\tilde F:(M,E)\to (M,E)$ that fixes E pointwise such that for every $p\in E$ the germ of $\tilde F$ at p is reduced according to the previous definition.

Model (i) corresponds to the points $p\in E$ which are not singular, in Abate’s terminology [Reference Abate1] (that is, the points which are non-singular for the saturation of the associated vector field), and models (ii) and (iii) correspond to the points $p\in E$ which are singular; for the latter, we use the names ‘non-degenerate’ and ‘saddle-node’ by analogy with the standard terminology for vector fields, according to whether the linear part of the saturation of the associated vector field has two or one non-zero eigenvalues, respectively. The dynamics of biholomorphisms of the form (i) is described in [Reference Abate1, Proposition 2.1] when $M=0$ and in [Reference Brochero Martínez3, Theorem 5.3] when $N=0$ .

In this paper we study the dynamics of non-degenerate fixed points. These are the only models that appear at singular points in the resolution of a generic biholomorphism.

Actually, we consider in our study a slightly more general class of biholomorphisms, since we do not impose the non-resonance condition $a/b\not \in \mathbb Q_{>0}$ . We distinguish two cases, according to whether the fixed point set of F has one or two components. As a first case, we consider biholomorphisms of the form

(A) $$ \begin{align} F(x,y)=(x+x^{M+1}[a+A(x,y)], y+x^M[cx+by+B(x,y)]), \end{align} $$

where $M\ge 1$ , $a,b,c\in \mathbb {C}$ , $ab\neq 0$ , $\operatorname {\mathrm {ord}} A\ge 1$ and $\operatorname {\mathrm {ord}} B\ge 2$ . Non-degenerate fixed points with $N=0$ are included here. Biholomorphisms of this type appear, for instance, after one blow-up at the point corresponding to a so-called non-degenerate characteristic direction (see [Reference Hakim8]). Écalle [Reference Écalle5] and Hakim [Reference Hakim8] showed that in this case there exist one-dimensional stable manifolds for F with 0 in their boundary, called parabolic curves. Moreover, if a and b satisfy the condition

(A) $$\begin{align} \operatorname{\mathrm{Re}}(b/a)>0, \end{align} $$

Hakim proved in [Reference Hakim7] (see also [Reference Arizzi and Raissy2]) the existence of two-dimensional stable manifolds with 0 in their boundary, called parabolic domains, where F is analytically conjugate to the map $(z,w)\mapsto (z+1,w)$ .

As a second case, we consider biholomorphisms of the form

(B) $$ \begin{align} F(x,y)=(x+x^{M+1}y^N[a+A(x,y)], y+x^My^{N+1}[b+B(x,y)]), \end{align} $$

where $M, N\ge 1$ , $ab\neq 0$ and $\operatorname {\mathrm {ord}} A,\operatorname {\mathrm {ord}} B\ge 1$ . Non-degenerate fixed points with $N\ge 1$ are included here. If a and b are such that $aM+bN\neq 0$ and satisfy the condition

(B) $$\begin{align} \operatorname{\mathrm{Re}}\bigg(\frac a{aM+bN}\bigg)>0\quad\text{and}\quad \operatorname{\mathrm{Re}}\bigg(\frac b{aM+bN}\bigg)>0, \end{align} $$

Vivas proved in [Reference Vivas15] the existence of parabolic domains for F.

The main result of this paper is an analog of the Leau–Fatou flower theorem for biholomorphisms of this type, providing a complete description of the dynamics in a whole neighborhood of the origin.

Theorem 1.2. Let F be a local biholomorphism of the form (A) or (B). In the first case, assume that F satisfies condition (A ) and set $d=M$ and $N=0$ ; in the second case, assume that F satisfies condition (B ) and set $d=\gcd (M,N)$ . Then, in any neighborhood of the origin there exist d pairwise disjoint connected open sets $\Omega ^+_0,\Omega ^+_1,\ldots ,\Omega ^+_{d-1}$ , with $0\in \partial \Omega ^+_k$ for all k, and d pairwise disjoint connected open sets $\Omega ^-_0,\Omega ^-_1,\ldots ,\Omega ^-_{d-1}$ , with $0\in \partial \Omega ^-_k$ for all k, such that the following assertions hold.

  1. (1) The sets $\Omega ^+_k$ are invariant for F and $F^j\to 0$ as $j\to +\infty $ compactly on $\Omega ^+_k$ for all k, and the sets $\Omega ^-_k$ are invariant for $F^{-1}$ and $F^{-j}\to 0$ as $j\to +\infty $ compactly on $\Omega ^-_k$ for all k.

  2. (2) The sets $\Omega ^+_0,\ldots ,\Omega ^+_{d-1},\Omega ^-_0,\ldots ,\Omega ^-_{d-1}$ together with the fixed set $\{xy^N=0\}$ cover a neighborhood of the origin.

  3. (3) For each k, there exist biholomorphisms $\varphi ^+_k: \Omega ^+_k\to W_k^+\subset \mathbb C^2$ and $\varphi ^-_k: \Omega ^-_k \to W_k^-\subset \mathbb C^2$ , with $W_k^+,W_k^-\subset \mathbb C\times \mathbb C^*$ if $N\ge 1$ , with the following properties:

    1. (a) $\varphi ^+_k$ and $\varphi ^-_k$ conjugate F with the map $(z,w)\mapsto (z+1,w)$ .

    2. (b) The sets $W_k^+$ and $W_k^-$ satisfy

      $$ \begin{align*}\bigcup_{\pm j\in\mathbb N} [W_k^{\pm}\kern1.4pt{-}\kern1.4pt(j,0)]\kern1.4pt{=}\kern1.4pt\mathbb C^2 \text{ if } N\kern1.4pt{=}\kern1.4pt0; \quad \bigcup_{\pm j\in\mathbb N} [W_k^{\pm}\kern1.4pt{-}\kern1.4pt(j,0)]\kern1.4pt{=}\kern1.4pt\mathbb C\kern1.5pt{\times}\kern1.5pt\mathbb C^* \text{ if } N\kern1.5pt{\ge}\kern1.5pt1.\end{align*} $$

Our second result shows that if conditions (A ) or (B ) are strictly not satisfied, then F has generic finite orbits in some neighborhood of the origin and so no two-dimensional stable sets. A version of this result has already been proved by Lisboa [Reference Lisboa10] for analytic vector fields.

Theorem 1.3. Let F be a local biholomorphism of the form (A) or (B). In the first case, assume that $\operatorname {\mathrm {Re}}(b/a)<0$ ; in the second case, assume that either

$$ \begin{align*}\operatorname{\mathrm{Re}}\bigg(\frac{a}{aM+bN}\bigg)<0\quad\text{or}\quad \operatorname{\mathrm{Re}}\bigg(\frac{b}{aM+bN}\bigg)<0.\end{align*} $$

Then there exists a neighborhood $\mathcal {U}$ of the origin and there exist sets $\mathcal {P}^+,\mathcal {P}^-\subset \mathcal {U}$ , which are one-dimensional submanifolds of $\mathcal U$ if F is of the form (A) and are empty otherwise, such that the following properties hold: given $p\in \mathcal {U}\setminus \mathcal P^+$ outside the fixed set, there exists $j\in \mathbb {N}$ such that $F^j(p)\notin \mathcal {U}$ ; given $p\in \mathcal {U}\setminus \mathcal P^-$ outside the fixed set, there exists $j\in \mathbb {N}$ such that $F^{-j}(p)\notin \mathcal {U}$ . If F is of the form (A), then $\mathcal P^+$ is the set of points in $\mathcal U$ that are attracted under F to the parabolic curves of F, and $\mathcal P^-$ is the set of points in $\mathcal U$ that are attracted under $F^{-1}$ to the parabolic curves of $F^{-1}$ .

The following example shows the necessity of the hypotheses in Theorems 1.2 and 1.3.

Example 1.4. For $M\ge 1$ , $N\ge 0$ and $a,b\in \mathbb {C}^*$ that do not satisfy the hypotheses in Theorems 1.2 and 1.3, consider the biholomorphism F given by the time-one flow of the vector field

$$ \begin{align*}X=x^My^N\bigg[ax{\dfrac{\partial}{\partial x}}+by{\dfrac{\partial}{\partial y}}\bigg]\end{align*} $$

and let us show that for any neighborhood $\mathcal {U}$ of the origin there exists $p\in \mathcal {U}$ outside the fixed set such that the orbit $\{F^j(p)\colon j\in \mathbb {Z}\}$ is contained in $\mathcal {U}$ and bounded away from the origin.

Assume first that $N=0$ , so $\operatorname {\mathrm {Re}}(b/a)=0$ . If $(x(t),y(t))$ is a solution of X and we set $(x_0,y_0)=(x(0),y(0))$ , we get by integration that

$$ \begin{align*}x(t)=x_0[1-aMx_0^Mt]^{-1/M}\quad\text{and}\quad y(t)=y_0[1-aMx_0^Mt]^{-b/(aM)},\end{align*} $$

defined for all $t\in \mathbb {R}$ provided that $aMx_0^M\notin \mathbb {R}$ . We have that $|x(t)|\le C|x_0|$ for some $C>0$ and for all $t\in \mathbb R$ , and if we set ${b}/{(aM)}=i\beta $ we have

$$ \begin{align*}|y(t)|=|y_0|e^{\beta\arg(1-aMx_0^Mt)},\end{align*} $$

so $e^{-|\beta |\pi }|y_0|\le |y(t)|\le e^{|\beta |\pi }|y_0|$ for all $t\in \mathbb R$ . Therefore, given a neighborhood $\mathcal {U}$ of the origin, if we choose $(x_0,y_0)$ as above and sufficiently small with $y_0\neq 0$ then its orbit is contained in $\mathcal {U}$ and bounded away from the origin.

Assume now that $N\ge 1$ , so either $aM+bN=0$ or $aM+bN\neq 0$ , $\operatorname {\mathrm {Re}}({a}/({aM+bN}))\ge 0$ and $\operatorname {\mathrm {Re}}({b}/({aM+bN}))=0$ . If $(x(t),y(t))$ is a solution of X with $(x(0),y(0))=(x_0,y_0)$ and we set $P(t)=x(t)^My(t)^N$ , we have $P'=(aM+bN)P^2$ , $x'=aPx$ and $y'=bPy$ . Suppose first that $aM+bN=0$ . Then $P(t)=P(0)$ , $x(t)=x_0e^{aP(0)t}$ and $y(t)=y_0e^{bP(0)t}$ , so

$$ \begin{align*}(x(t),y(t))=(x_0e^{ax_0^My_0^Nt},y_0e^{bx_0^My_0^Nt})\end{align*} $$

for all $t\in \mathbb R$ . Note that, since $a/b=-N/M\in \mathbb R$ , in any neighborhood $\mathcal {U}$ of the origin we can take $(x_0,y_0)$ arbitrarily small with $x_0y_0\neq 0$ such that $\operatorname {\mathrm {Re}} (ax_0^{M}y_0^N)=\operatorname {\mathrm {Re}} (bx_0^{M}y_0^N)=0$ . In this case, the expression above shows that $|x(t)|=|x_0|$ and $|y(t)|=|y_0|$ for all $t\in \mathbb R$ , so the orbit of $(x_0,y_0)$ is bounded away from the origin and contained in $\mathcal {U}$ provided $(x_0,y_0)$ is small enough. Suppose now that $aM+bN\neq 0$ , $\operatorname {\mathrm {Re}}({a}/({aM+bN}))\ge 0$ and $\operatorname {\mathrm {Re}}({b}/({aM+bN}))=0$ . By integration we have that $P(t)=x_0^My_0^N[1-(aM+bN)x_0^My_0^Nt]^{-1}$ and

$$ \begin{align*} x(t)&=x_0[1-(aM+bN)x_0^My_0^Nt]^{-a/(aM+bN)},\\ y(t)&=y_0[1-(aM+bN)x_0^My_0^Nt]^{-b/(aM+bN)}, \end{align*} $$

defined for all $t\in \mathbb {R}$ provided $(aM+bN){x_0^My_0^N}\notin \mathbb {R}$ . If we set ${a}/({aM+bN})=\alpha _1+i\alpha _2$ , with $\alpha _1\ge 0$ , and ${b}/({aM+bN})=i\beta $ , we have

$$ \begin{align*}|x(t)|=|x_0||1-(aM+bN)x_0^My_0^Nt|^{-\alpha_1}e^{\alpha_2\arg(1-(aM+bN)x_0^My_0^Nt)}\end{align*} $$

and

$$ \begin{align*}|y(t)|=|y_0|e^{\beta\arg(1-(aM+bN)x_0^My_0^Nt)},\end{align*} $$

so

$$ \begin{align*}|x(t)|\le C e^{|\alpha_2|\pi}|x_0|,\quad e^{-|\beta|\pi}|y_0|\le |y(t)|\le e^{|\beta|\pi}|y_0|\end{align*} $$

for some $C>0$ and for all $t\in \mathbb R$ . Therefore, given a neighborhood $\mathcal {U}$ of the origin, if we choose $(x_0,y_0)$ as above and sufficiently small then its orbit is contained in $\mathcal {U}$ and bounded away from the origin.

To conclude the introduction, let us briefly describe the structure of the paper. In §2 we prove some basic dynamic facts in the spirit of the Leau–Fatou flower theorem that we will use throughout the paper. Sections 37 are devoted to the proof of Theorem 1.2. In §§3 and 4 we show the existence of invariant domains and invariant functions for maps of the form (B), with $N\ge 0$ , satisfying the hypotheses of Theorem 1.2; as we will explain in §7, this will also allow us to obtain Theorem 1.2 for maps of the form (A). In §5 we construct an approximation of Fatou coordinates (that is, conjugations with $(z,w)\mapsto (z+1,w)$ ), which we modify to actual Fatou coordinates in §6. The final details of the proof of Theorem 1.2 are provided in §7. Finally, in §8 we prove Theorem 1.3.

2 Stable subdynamics in two variables

In this section we prove Proposition 2.2, which provides some basic dynamic facts in the spirit of the Leau–Fatou flower theorem, adapted to our two-dimensional context.

Definition 2.1. Given $d\in \mathbb N$ , $\varepsilon>0$ and $\theta \in (0,\pi /2)$ , we define the sets

$$ \begin{align*}S(d,\varepsilon,\theta)=\{z\in\mathbb C:|z^d|<\varepsilon, \lvert\arg(z^d)\rvert<\theta\}, \end{align*} $$

whose connected components are the d sectors

$$ \begin{align*}S_k(d,\varepsilon,\theta)= \bigg\{z\in\mathbb C:|z|<\varepsilon^{1/d}, \bigg\lvert\!\arg z-\dfrac{2k\pi }d\bigg\rvert<\frac \theta d\bigg\}\end{align*} $$

for $k=0,\ldots ,d-1$ , and

$$ \begin{align*}\widetilde S(d,\varepsilon,\theta)=S(d,\varepsilon,\theta)\cup \bigg\{z\in\mathbb C: \bigg|z^d-\dfrac{\varepsilon}2 e^{-i\theta}\bigg|<\dfrac{\varepsilon}2\bigg\}\cup \bigg\{z\in\mathbb C:\bigg|z^d-\dfrac{\varepsilon}2 e^{i\theta}\bigg|<\dfrac{\varepsilon}2\bigg\},\end{align*} $$

whose connected components are d sectorial domains $\widetilde S_k(d,\varepsilon ,\theta )$ of opening $(\pi +2\theta )/d$ bisected by the rays $e^{{2k\pi i }/{d}}\mathbb R^+$ for $k=0,\ldots ,d-1$ .

Proposition 2.2. Let $G:\Omega \to \mathbb C^2$ be a holomorphic function, where $\Omega $ is a neighborhood of $0\in \mathbb C^2$ . Denote $(x_j,y_j)=G^j(x,y)$ for all $j\ge 0$ . Let $m,n,d\in \mathbb Z_{\ge 0}$ be such that $m+n\ge 1$ and $d\ge 1$ and suppose that

$$ \begin{align*} x_1^my_1^n=x^my^n-\frac1d(x^my^n)^{d+1}+(x^my^n)^{d+1}O(x,y). \end{align*} $$

There exist functions $\varepsilon _G, \delta _G , \tilde \delta _G:(0,\pi /2)\to \mathbb R^+$ with the following properties.

  1. (1) If $\mathcal S$ is a connected component of $S(d,\varepsilon ,\theta )$ or $\widetilde S(d,\varepsilon ,\theta )$ with $\theta \in (0,\pi /2)$ and $\varepsilon \le \varepsilon _G(\theta )$ , then for all $(x,y)$ with $x^my^n\in \mathcal S$ , $|x|<\delta _G({\theta })$ and $|y|<\delta _G({\theta })$ we have that $x_1^my_1^n\in \mathcal S$ . Hence, if $x^my^n\in \mathcal S$ and $|x_j|<\delta _G(\theta )$ , $|y_j|<\delta _G(\theta )$ for all $j\ge 0$ then $x_j^my_j^n\in \mathcal S $ for all $j\ge 0$ . In this case, if $\mathcal S$ is a component of $S(d,\varepsilon ,\theta )$ we have that

    $$ \begin{align*} \frac12\frac{|x^my^n|^d}{1+j|x^my^n|^d}\le |x_j^my_j^n|^d\le \frac 2{\cos\theta}\frac{|x^my^n|^d}{1+j|x^my^n|^d}\ \end{align*} $$
    for all $j\ge 0$ , and if $\mathcal S$ is a component of $\widetilde S(d,\varepsilon ,\theta )$ we have that
    $$ \begin{align*} &|x_j^my_j^n|^d\ge \frac 12 \frac{|x^my^n|^d}{1+j|x^my^n|^d}\quad\text{for all } j\ge 0, \\ &|x_j^my_j^n|^d\le 2\frac{|x^my^n|^d}{1+j|x^my^n|^d} \quad\text{for all } j\ge \frac 6{|x^my^n|^d}. \end{align*} $$

    In particular, $x_j^my_j^n\to 0$ as $j\to +\infty $ .

  2. (2) Consider a component $\widetilde S_k(d,\varepsilon ,\theta )$ of $\widetilde S(d,\varepsilon ,\theta )$ with $\varepsilon \le \varepsilon _G(\theta )$ and assume the following additional hypothesis on G: there exist $\nu>0$ and $\delta _\nu>0$ such that

    $$ \begin{align*}|x_1|\le|x|(1+\nu|x^my^n|^d)\quad\text{and}\quad |y_1|\le|y|(1+\nu|x^my^n|^d)\end{align*} $$
    whenever $x^my^n\in \widetilde S_k(d,\varepsilon ,\theta )$ , $|x|<\delta _\nu $ and $|y|<\delta _\nu $ . Then, given ${\mu>0}$ , we find ${\tilde \mu>0}$ with the following property: if $x^my^n\in \widetilde S_k(d,\varepsilon ,\theta )$ , $|x|<\tilde \mu $ and $|y|<\tilde \mu $ , there exists $j_0\ge 0$ such that $x^m_{j_0}y^n_{j_0}\in S_k(d,\varepsilon ,\theta )$ and, for every $j\le j_0$ , $ x^m_{j}y^n_{j}\in \widetilde S_k(d,\varepsilon ,\theta )$ , $|x_j|<\mu $ and $|y_j|<\mu $ .
  3. (3) Take $\theta \in (0,\pi /2)$ and consider a point $(x,y)$ with $x^my^n\neq 0$ such that $|x_j|<\tilde \delta _G({\theta })$ and $|y_j|<\tilde \delta _G(\theta )$ for all $j\ge 0$ . Then $x_j^my_j^n\to 0$ as $j\to +\infty $ and there exists $k\in \{0,\ldots , d-1\}$ such that $x_j^my_j^n\in S_k(d,\varepsilon ,\theta )$ for any $\varepsilon>0$ and for $j\ge 0$ big enough. Moreover, the two last inequalities of assertion 1 hold.

We will need the following result.

Lemma 2.3. There exists $\kappa \colon (0,\pi /2)\to \mathbb R^+$ with the following property. If ${f(z)= z-z^2}$ for $z\in \mathbb C$ and either $S=S(1,\varepsilon ,\theta )$ or $S=\widetilde S(1,\varepsilon ,\theta )$ with $\theta \in (0,\pi /2)$ and $\varepsilon \le {\kappa (\theta )}$ , then $ f(S)\subset S$ and for all $z\in S$ ,

$$ \begin{align*}\operatorname{\mathrm{dist}} (f(z),\partial S)\ge \kappa (\theta) |z|^2. \end{align*} $$

Proof. Suppose first that $S= S(1,\varepsilon ,\theta )$ . Set $\kappa (\theta )=\frac 18 \sin (\theta /2)$ and assume that $\varepsilon \le {\kappa (\theta )}$ . An easy computation shows that the map $z\mapsto w= 1 /z$ conjugates f to

$$ \begin{align*}g(w)= w+1+\frac {1}{w-1}\end{align*} $$

and transforms S into

$$ \begin{align*}W=\{w\in\mathbb C\colon |w|>1/\varepsilon,\ \lvert\arg w\rvert<\theta\}.\end{align*} $$

Fix $z\in S(1,\varepsilon ,\theta )$ and set $z_1=f(z)$ , $w=1/z$ and $w_1=1/z_1=g(w)$ . Consider the sectors $S_w(\theta )$ and $S_w(\theta /2)$ , where

$$ \begin{align*}S_w(\alpha)=w+\{\zeta\in\mathbb C \colon |\zeta|>0,\ \lvert\arg\zeta\rvert<\alpha\}.\end{align*} $$

Clearly, $S_w(\theta /2)\subset S_w(\theta )\subset W$ and

$$ \begin{align*}\operatorname{\mathrm{dist}} (w+1, \partial S_w(\theta/2))=\sin(\theta/ 2).\end{align*} $$

Then, since

$$ \begin{align*}\operatorname{\mathrm{dist}} (w+1, w_1)=\frac{1}{|w-1|}<\frac{1}{1/{\kappa(\theta)}-1} =\frac 1{8/\sin(\theta/2)-1}<\frac 1 4\sin(\theta/2)<\sin(\theta/2),\end{align*} $$

we see that $w_1\in S_w(\theta /2)\subset W$ , which proves that $f(S)\subset S$ . Note that

$$ \begin{align*}\operatorname{\mathrm{dist}} (w_1, \partial S_w(\theta))=|w_1-w|\sin\beta,\end{align*} $$

where $\beta $ is one of the angles determined by the ray $\overrightarrow {ww_1}$ in the sector $S_w(\theta )$ . Since ${w_1\in S_w(\theta /2)}$ , we have that $\beta \ge \theta /2$ and therefore

$$ \begin{align*} \operatorname{\mathrm{dist}} (w_1, \partial S_w(\theta))&>|w_1-w|\sin(\theta/2)\ge\bigg(1- \frac{1}{|w-1|}\bigg)\sin(\theta/2)\\ &> \big(1- \tfrac 1 4\sin(\theta/2)\big) \sin(\theta/2) >\tfrac 3 4 \sin(\theta/2). \end{align*} $$

Then, since $S_w(\theta )\subset W$ ,

$$ \begin{align*} \operatorname{\mathrm{dist}} (w_1, \partial W)\ge \operatorname{\mathrm{dist}} (w_1, \partial S_w(\theta))>\frac 34 \sin(\theta/2), \end{align*} $$

so the disk $\mathcal D=\{\zeta \in \mathbb C\colon |\zeta -w_1|<\frac 34 \sin (\theta /2)\}$ is contained in W. If we set $h(\zeta )=1/\zeta $ , we have that $h(W)=S$ and $h(\mathcal D)$ contains the disk $\{\xi \in \mathbb C\colon |\xi -h(w_1)|<r\}$ , where

$$ \begin{align*}r=\min\{|h(\zeta)-h(w_1)|\colon \zeta\in\mathcal \partial \mathcal D \}.\end{align*} $$

Notice that if $\zeta \in \mathcal \partial \mathcal D$ , then $|\zeta |\le |w_1|+3/4<2|w_1|$ . Then

$$ \begin{align*}r=\min\bigg\{\bigg|\frac{\zeta-w_1}{\zeta w_1}\bigg|\colon \zeta\in\mathcal \partial \mathcal D\bigg \}> \frac{ \frac{3}{4} \sin(\theta/2)}{2|w_1|^2}= \frac{ \frac{3}{8} \sin(\theta/2)}{|w_1|^2}. \end{align*} $$

Thus, since

$$ \begin{align*}1/{|w_1|^2}=|z_1|^2\ge |z|^2(1-|z|)^2\ge |z|^2(1-\kappa(\theta))^2\ge|z|^2 (1-1/8)^2>\tfrac 23 |z|^2,\end{align*} $$

we see that $r> \frac 14 \sin (\theta /2)|z|^2>\kappa (\theta ) |z|^2$ . Therefore we conclude that S contains the disk $\{\zeta \in \mathbb C\colon |\zeta -z_1|<\kappa (\theta ) |z|^2\}, $ which concludes the proof.

Suppose now that $S= \widetilde S(1,\varepsilon ,\theta )$ , take $\kappa (\theta )=\frac 18 \sin (\pi /4-\theta /2)$ and assume that ${\varepsilon \le {\kappa (\theta )}}$ . Then the map $z\mapsto w=1/z$ transforms S into

$$ \begin{align*}\widetilde W=W\cup\{w\in\mathbb C\colon\operatorname{\mathrm{Re}}(we^{-i\theta})>1/\varepsilon\}\cup\{w\in\mathbb C\colon \operatorname{\mathrm{Re}}(we^{i\theta})>1/\varepsilon\}.\end{align*} $$

Note that $S_w(\pi /4-\theta /2)\subset S_w(\pi /2-\theta )\subset W$ . Then the proof follows exactly as in previous case, with $\pi /2-\theta $ instead of $\theta $ . Finally, to unify the selection of $\kappa $ we can take $\kappa (\theta )= \frac 18\min \{\sin (\theta /2), \sin (\pi /4-\theta /2)\}$ , which works for both cases.

Proof of Proposition 2.2

From the expression for $x_1^my_1^n$ we obtain that

(1) $$ \begin{align} (x_1^my_1^n)^d=(x^my^n)^d-(x^my^n)^{2d}+(x^my^n)^{2d}\sigma(x,y) \end{align} $$

and

(2) $$ \begin{align} \frac 1{(x_1^my_1^n)^d}=\frac 1{(x^my^n)^d}+ 1+\sigma_1(x,y), \end{align} $$

where $\sigma =O(x,y)$ and $\sigma _1=O(x,y)$ . Moreover, notice that $x_1^my_1^n/(x^my^n)=1+ O(x,y)$ is arbitrarily close to 1 provided x and y are small enough. Then, given $\theta \in (0,\pi /2)$ , we can find $\delta _G(\theta )>0$ such that for all $(x,y)$ with $|x|<\delta _G(\theta )$ and $|y|<\delta _G(\theta )$ we have that

$$ \begin{align*} |\sigma(x,y)|< \kappa (\theta), |\sigma_1(x,y)|<\frac 14 \quad \text{and} \quad \bigg\lvert\!\arg\frac{x_1^my_1^n}{x^my^n}\bigg\rvert<\frac{\pi-2\theta}d, \end{align*} $$

where $\kappa (\theta )$ is given by Lemma 2.3. Set $\varepsilon _G(\theta )= {\kappa (\theta )}.$

Let us prove assertion 1 of the proposition. Take $\varepsilon \le \varepsilon _G(\theta )$ , let $\mathcal S$ be one of the components of $S(d,\varepsilon ,\theta )$ or $\widetilde S(d,\varepsilon ,\theta )$ and take $(x,y)$ such that $x^my^n\in \mathcal S$ , $|x|<\delta _G(\theta )$ and $|y|<\delta _G(\theta )$ . Set $f(z)=z-z^2$ . Note that $S=\{z^d\colon z\in \mathcal S\}$ is one of the sets $S(1,\varepsilon ,\theta )$ or $\widetilde S(1,\varepsilon ,\theta )$ , so by equation (1) and Lemma 2.3 we have that

$$ \begin{align*} |(x_1^my_1^n)^d-f((x^my^n)^d)|<\kappa (\theta) |x^my^n|^{2d}\le \operatorname{\mathrm{dist}} (f((x^my^n)^d),\partial S), \end{align*} $$

which means that $(x_1^my_1^n)^d\in S$ . Since the components of $S(d,\varepsilon ,\theta )$ or $\widetilde S(d,\varepsilon ,\theta )$ are separated by a sector of opening at least $(\pi -2\theta )/d$ and $\lvert \arg (x_1^my_1^n/(x^my^n))\rvert < (\pi -2\theta )/d$ , we conclude that $(x_1^my_1^n)^d$ belongs to the same component $\mathcal {S}$ .

Now, let $(x,y)$ be such that $x^my^n\in \mathcal S$ with $|x_j|<\delta _G(\theta )$ and $|y_j|<\delta _G(\theta )$ for all $j\ge 0$ . From (2) we obtain that

$$ \begin{align*} \frac 1{(x_j^my_j^n)^d}=\frac 1{(x^my^n)^d}+ j+\sum_{l=0}^{j-1}\sigma_1(x_l,y_l), \end{align*} $$

so

$$ \begin{align*} \frac 1{|x_j^my_j^n|^d}\le \frac 1{|x^my^n|^d}+j+ \frac 14 j \le 2\bigg(\frac 1{|x^my^n|^d}+j\bigg) \end{align*} $$

for all $j\ge 0$ , which gives the lower bounds in the inequalities of assertion 1. On the other hand, from the equation above we have that

$$ \begin{align*} \frac 1{|x_j^my_j^n|^d} &\ge \operatorname{\mathrm{Re}} \frac 1{(x_j^my_j^n)^d} \ge \operatorname{\mathrm{Re}} \frac 1{(x^my^n)^d}+j- \sum_{l=0}^{j-1}|\sigma_1(x_l,y_l)|\\ &\ge \operatorname{\mathrm{Re}} \frac 1{(x^my^n)^d}+j - \frac 14j = \operatorname{\mathrm{Re}} \frac 1{(x^my^n)^d}+ \frac 34 j, \end{align*} $$

for all $j\ge 0$ . If $\mathcal S$ is a component of $S(d,\varepsilon ,\theta )$ , since $\lvert \arg ( 1/ {(x^my^n)^d}\rvert <\theta $ , we have that $\operatorname {\mathrm {Re}} ( 1/{(x^my^n)^d})\ge {\cos \theta }/{|x^my^n|^d}$ and therefore

$$ \begin{align*} \frac 1{|x_j^my_j^n|^d}\ge \frac {\cos \theta}{|x^my^n|^d}+ \frac 34 j \ge \frac{\cos\theta}2\bigg(\frac 1{|x^my^n|^d}+j \bigg) \end{align*} $$

for all $j\ge 0$ , which gives the upper bound in the first inequality of assertion 1. If $\mathcal S$ is a component of $\widetilde S(d,\varepsilon ,\theta )$ we obtain that for all $j\ge 6/|x^my^n|^d$ ,

$$ \begin{align*} \frac 1{|x_j^my_j^n|^d} \ge \operatorname{\mathrm{Re}}\frac 1{(x^my^n)^d}+\frac 34 j \ge -\frac 1{|x^my^n|^d}+ \frac 34 j \ge \frac 12 \bigg(\frac 1{|x^my^n|^d} +j\bigg), \end{align*} $$

which concludes the proof of assertion 1.

Let us prove assertion 2. From the expression for $x_1^my_1^n$ we can write $x_1^my_1^n=x^my^n(1-\zeta ),$ where $\zeta =(1/d+\tau ) (x^my^n)^d$ and $\tau =O(x,y)$ . Then, up to reducing $\delta _\nu $ , we have that if $ |x|<\delta _\nu $ and $ |y|<\delta _\nu $ then

$$ \begin{align*} |\zeta|<1,\quad \lvert\arg(1-\zeta)\rvert<\theta/d,\quad |1/d+\tau|>\delta_\nu, \quad \lvert\arg(1/d+\tau)\rvert<\theta^*, \end{align*} $$

where $\theta ^*= \min \{\theta /2, \pi /4-\theta /2\}$ . Set

$$ \begin{align*}\tilde\mu=e^{-2\pi\nu/(d\delta_\nu\sin\theta^*)}\min\{\mu,\delta_\nu, \delta_G\}.\end{align*} $$

Take $x^my^n\in \widetilde S_k(d,\varepsilon ,\theta )$ such that $|x|<\tilde \mu $ and $|y|<\tilde \mu $ . If $\lvert \arg (x^my^n)^d\rvert <\theta $ , we have that $x^my^n\in S_k(d,\varepsilon ,\theta )$ and, since $\tilde \mu \le \mu $ , we can take $j_0=0$ and we are done. Thus we can assume that $\lvert \arg (x^my^n)^d\rvert \in [\theta , \theta +\pi /2)$ . Moreover, we assume that $\arg (x^my^n)^d\in [\theta , \theta +\pi /2)$ ; the other case is analogous. Then

$$ \begin{align*}\arg\zeta=\arg(1/d+\tau)+\arg (x^my^n)^d\in (\theta-\theta^*, \theta+\pi/2+\theta^*)\subset (\theta/2, \theta/2+3\pi/4)\end{align*} $$

and hence

$$ \begin{align*}\operatorname{\mathrm{Im}}\zeta>|\zeta|\min\{\sin(\theta/2),\sin(\theta/2+3\pi/4)\}=|\zeta|\sin\theta^*.\end{align*} $$

Then, since $|\zeta |<1$ , $\sin (\arg (1-\zeta ))=- {\operatorname {\mathrm {Im}}\zeta }/{|1-\zeta |} < -\tfrac 1 2 \sin \theta ^*|\zeta |$ and therefore

$$ \begin{align*}-\theta <\arg (1-\zeta)^d<d\sin(\arg (1-\zeta))< -\frac d2\sin\theta^*|\zeta|.\end{align*} $$

It follows that

$$ \begin{align*}0\le\arg(x^my^n)^d- \theta &< \arg(x^my^n)^d+\arg(1-\zeta)^d < \arg(x^my^n)^d-\frac d2 \sin\theta^*|\zeta|\\ &<\arg(x^my^n)^d-\frac d2 \delta_\nu{ \sin\theta^*} |x^my^n|^d; \end{align*} $$

that is, $\arg (x_1^my_1^n)^d=\arg (x^my^n)^d+\arg (1-\zeta )^d$ satisfies

(3) $$ \begin{align} 0< \arg(x_1^my_1^n)^d< \arg(x^my^n)^d-\frac d2 \delta_\nu \sin\theta^* |x^my^n|^d. \end{align} $$

Then $ |x^my^n|^d<(2/(d\delta_{\nu}\sin\theta^*))\arg (x^my^n)^d< 2\pi/(d\delta_{\nu}\sin\theta^*)$ so

$$ \begin{align*}|x_1|\le |x|(1+\nu |x^my^n|^d)<|x| e^{\nu |x^my^n|^d}<\tilde\mu e^{ 2\pi\nu/(d\delta_\nu\sin\theta^*)}=\min\{\mu,\delta_\nu,\delta_G\}\end{align*} $$

and, analogously, $|y_1|<\min \{\mu , \delta _\nu ,\delta _G\}$ . Thus, if $\arg (x_1^my_1^n)^d<\theta $ , since $x_1^my_1^n\in \widetilde S_k(d,\varepsilon , \theta )$ because of assertion 1, we have that $x_1^my_1^n\in S_k(d,\varepsilon ,\theta )$ , we can choose $j_0=1$ and we are done. We assume then that $\arg (x_1^my_1^n)^d\in [\theta , \theta +\pi /2)$ . Proceeding exactly as above, we obtain that

$$ \begin{align*}0 <\arg(x_2^my_2^n)^d < \arg(x_1^my_1^n)^d-\frac d2 \delta_\nu{ \sin\theta^*} |x_1^my_1^n|^d \end{align*} $$

and, in view of (3),

$$ \begin{align*} 0< \arg(x_2^my_2^n)^d <\arg(x^my^n)^d-\frac d2 \delta_\nu{ \sin\theta^*} |x^my^n|^d-\frac d2 \delta_\nu{ \sin\theta^*} |x_1^my_1^n|^d. \end{align*} $$

Then

$$ \begin{align*}|x^my^n|^d+ |x_1^my_1^n|^d<\frac 2 {d\delta_\nu{ \sin\theta^*} }\arg(x^my^n)^d<\frac {2\pi} {d\delta_\nu{ \sin\theta^*} }\end{align*} $$

so

$$ \begin{align*} |x_2|&\le |x|(1+\nu |x^my^n|^d)(1+\nu |x_1^my_1^n|^d)<|x| e^{\nu |x^my^n|^d+\nu |x_1^my_1^n|^d}\\ &<\tilde\mu e^{2\pi\nu/(d\delta_\nu\sin\theta^*)}=\min\{\mu,\delta_\nu, \delta_G\} \end{align*} $$

and, analogously, $|y_2|<\min \{\mu ,\delta _\nu ,\delta _G\}$ . Thus, if $\arg (x_2^my_2^n)^d<\theta $ , we have that $x_2^my_2^n\in S_k(d,\varepsilon ,\theta )$ , we can take $j_0=2$ and we are done. We assume then that $\arg (x_2^my_2^n)^d\in [\theta , \theta +\pi /2)$ and repeat the argument. If assertion 2 were false, this process would continue indefinitely and we would obtain, for every $j\ge 0$ , that $|x_j|<\delta _G$ , $|y_j|<\delta _G$ and

$$ \begin{align*} |x^my^n|^d+ |x_1^my_1^n|^d+\cdots + |x_j^my_j^n|^d<\frac {2\pi} {d\delta_\nu{ \sin\theta^*} }.\end{align*} $$

But the inequalities in assertion 1 show that $\sum _{j\ge 0}|x_j^my_j^n|^d$ diverges, which is a contradiction.

Let us prove assertion 3. Let $\theta \in (0,\pi /2)$ . Choose $\delta _1>0$ such that for all $(x,y)$ with $|x|<\delta _1$ and $|y|<\delta _1$ we have $|\sigma _1(x,y)|<\tfrac 12 \tan \theta $ and set

$$ \begin{align*}\tilde\delta_G(\theta)=\min\{\delta_1, \delta_G,\varepsilon_G ^{1/((m+n)d)}\}.\end{align*} $$

Take $(x,y)$ such that $|x_j|<\tilde \delta _G(\theta )$ and $|y_j|<\tilde \delta _G(\theta )$ for all $j\ge 0$ . From (2) we have

$$ \begin{align*} \bigg\vert\!\operatorname{\mathrm{Im}} \frac 1{(x_j^my_j^n)^d}\bigg\vert\le \bigg\vert\!\operatorname{\mathrm{Im}}\frac 1{(x^my^n)^d}\bigg|+ \sum_{l=0}^{j-1}|\sigma_1(x_l,y_l)| \le \bigg\vert\!\operatorname{\mathrm{Im}}\frac 1{(x^my^n)^d}\bigg|+\frac 12 j\tan\theta. \end{align*} $$

On the other hand, as we showed in the proof of assertion 1,

$$ \begin{align*} \operatorname{\mathrm{Re}} \frac 1{(x_j^my_j^n)^d} \ge \operatorname{\mathrm{Re}}\frac 1{(x^my^n)^d}+\frac 34 j. \end{align*} $$

Therefore, for j big enough,

$$ \begin{align*} 0< \frac {\lvert\operatorname{\mathrm{Im}} ( 1/{(x_j^my_j^n)^d})\rvert}{\operatorname{\mathrm{Re}} {( 1/{(x_j^my_j^n)^d})}}\le \frac {\lvert\operatorname{\mathrm{Im}}( 1/{(x^my^n)^d})\rvert+ (1/2) j\tan \theta} { \operatorname{\mathrm{Re}}( 1/{(x^my^n)^d})+(3/4) j}< \tan\theta. \end{align*} $$

This means that $\lvert \arg ( 1/{(x_j^my_j^n)^d})\rvert <\theta $ for j big enough, so we can fix $j_0\ge 0$ such that $\lvert \arg (x_{j_0}^my_{j_0}^n)^d\rvert <\theta $ . Moreover, since $|(x_{j_0}^my_{j_0}^n)^d|<\tilde \delta _G(\theta ) ^{(m+n)d}\le \varepsilon _G (\theta )$ , we have that $x_{j_0}^my_{j_0}^n\in S(d,\varepsilon _G (\theta ),\theta )$ . Thus $x_{j_0}^my_{j_0}^n\in S_k(d,\varepsilon _G (\theta ),\theta )$ for some $k\in \{0,\ldots , d-1\}$ and it follows from assertion 1 that $x_{j}^my_{j}^n\in S_k(d,\varepsilon _G (\theta ),\theta )$ for all $j\ge j_0$ and $x_{j}^my_{j}^n\to 0$ . Clearly, given $\varepsilon>0$ , up to increasing $j_0$ , we have that $x_{j}^my_{j}^n\in S_k(d,\varepsilon ,\theta )$ for all $j\ge j_0$ . Moreover, arguing exactly as in the proof of assertion 1 we have that the two last inequalities of that assertion hold. This concludes the proof of Proposition 2.2.

3 Existence of parabolic domains

In this section we show the existence of parabolic domains for biholomorphisms of the form (B) satisfying condition (B ). Actually, we consider a slightly larger class of biholomorphisms, since we allow $N\ge 0$ (note that if $N=0$ condition (B ) is precisely condition (A )).

Consider a biholomorphism of the form (B), with $N\ge 0$ , satisfying condition (B ). Set $d=M$ if $N=0$ and $d=\gcd (M,N)$ otherwise. Applying a linear change of coordinates of the form $(x,y)\mapsto (\alpha x,\beta y)$ , we obtain the same expression for F but with a and b respectively replaced by $\tilde a=-{a}/({aM+bN})$ and $ \tilde b=-{b}/({aM+bN}),$ so hypothesis (B ) becomes $\operatorname {\mathrm {Re}} \tilde {a}<0$ and $\operatorname {\mathrm {Re}} \tilde {b}<0$ , and we have $\tilde {a}M+\tilde {b}N=-1$ . Thus, we directly assume that

$$ \begin{align*}\operatorname{\mathrm{Re}} a<0, \quad \operatorname{\mathrm{Re}} b<0 \quad \text{and} \quad aM+bN=-1.\end{align*} $$

Set $m={M}/{d}$ and $n={N}/{d}$ . Given a point $(x,y)$ in the domain of definition of F, we denote $(x_j,y_j)=F^j(x,y)$ for all $j\ge 0$ . Notice that, from the expression for F, we easily obtain that

(4) $$ \begin{align} x_1^my_1^n=x^my^n-\frac 1d(x^my^n)^{d+1}+(x^my^n)^{d+1}O(x,y). \end{align} $$

Definition 3.1. Fix $\gamma \in (0,1)$ such that

$$ \begin{align*} \gamma (m+n)\le1,\quad \operatorname{\mathrm{Re}} \bigg(a+\dfrac\gamma d\bigg)<0\quad \text{and} \quad \operatorname{\mathrm{Re}}\bigg( b+\dfrac\gamma d\bigg)<0 \end{align*} $$

and set $\iota =0$ if $n=0$ and $\iota =1$ if $n\ge 1$ . Given $\theta \in (0, {\pi }/2)$ and $\varepsilon , \delta \in (0,1]$ , we consider the sets $D_k=D_k(\varepsilon ,\theta ,\delta )$ , $\widetilde D_k=\widetilde D_k(\varepsilon ,\theta ,\delta )$ and $U_k=U_k(\varepsilon ,\theta )$ , for $k\in \{0,\ldots ,d-1\}$ , defined by

$$ \begin{align*}D_k=\{(x,y)\in\mathbb C^2\colon x^my^n\in S_k(d,\varepsilon,\theta), |\iota x|<\delta, |y|<\delta\},\end{align*} $$
$$ \begin{align*}\widetilde D_k=\{(x,y)\in\mathbb C^2\colon x^my^n\in \widetilde S_k(d,\varepsilon,\theta), |\iota x|<\delta, |y|<\delta\},\end{align*} $$

where $S_k(d,\varepsilon ,\theta )$ and $\widetilde S_k(d,\varepsilon ,\theta )$ are the sets of Definition 2.1, and

$$ \begin{align*}U_k=\{(x,y)\in\mathbb C^2\colon x^my^n\in S_k(d,\varepsilon,\theta), |\iota x|\le|x^my^n|^\gamma, |y|\le|x^my^n|^\gamma\}.\end{align*} $$

Remark 3.2. If $n\ge 1$ , the condition $x^my^n\in S_k(d,\varepsilon ,\theta )$ or $x^my^n\in \widetilde S_k(d,\varepsilon ,\theta )$ implies that the sets $D_k$ , $\widetilde D_k$ and $U_k$ are disjoint from the coordinate axes $\{xy=0\}$ . If $n=0$ , we have that $x^my^n=x$ so the sets $D_k$ , $\widetilde D_k$ and $U_k$ are disjoint from $\{x=0\}$ but they intersect $\{y=0\}$ . On the other hand, if $n=0$ notice that $U_k=\{x\in S_k(d,\varepsilon ,\theta ), |y|\le |x|^\gamma \}$ and, since $\varepsilon ,\gamma \le 1$ , the inequality $|x|\le |x|^\gamma $ holds in $U_k$ . Therefore, even if $n=0$ we can write

$$ \begin{align*}U_k=\{(x,y)\in\mathbb C^2:x^my^n\in S_k(d,\varepsilon,\theta), |x|\le|x^my^n|^\gamma, |y|\le|x^my^n|^\gamma\}.\end{align*} $$

Finally note that, because of the condition $\gamma (m+n)\le 1$ , the set $U_k$ is non-empty.

Since $\lvert \arg (-a)\rvert <\pi /2$ and $\lvert \arg (-b)\rvert <\pi /2$ and, by the choice of $\gamma $ , $\lvert \arg (-a-\gamma /d)\rvert <\pi /2$ and $\lvert \arg (-b-\gamma /d)\rvert <\pi /2$ , we can fix $\theta _0\in (0,\pi /2)$ such that

$$ \begin{align*} \begin{aligned} \lvert\arg(- a)\rvert+2\theta_0<\pi/2,\quad \lvert\arg(-a-\gamma/d)\rvert+2\theta_0<\pi/2, \\ \lvert\arg(- b)\rvert+2\theta_0<\pi/2 ,\quad \lvert\arg(-b-\gamma/d)\rvert+2\theta_0<\pi/2. \end{aligned} \end{align*} $$

Proposition 3.3. Let F be a biholomorphism of the form (B) with $N\ge 0$ satisfying condition (B ) and fix $\theta _0$ as above. Given $\theta \in (0,\theta _0)$ , there exist $\varepsilon _\theta , \delta _\theta \in (0,1]$ such that for $\varepsilon \le \varepsilon _\theta $ and $\delta \le \delta _\theta $ the following properties hold.

  1. (1) For every $k\in \{0,\ldots , d-1\}$ we have that

    $$ \begin{align*}F(D_k(\varepsilon,\theta,\delta))\subset D_k(\varepsilon,\theta,\delta), \quad F(U_k(\varepsilon,\theta))\subset U_k(\varepsilon,\theta),\end{align*} $$
    $F^j\to 0$ as $j\to +\infty $ uniformly on $U_k(\varepsilon , \theta )$ , $D_k(\varepsilon ,\theta ,\delta )$ is contained in the basin of attraction of $U_k(\varepsilon ,\theta )$ , and every point $(x,y)\in U_k(\varepsilon ,\theta )\cup D_k(\varepsilon ,\theta ,\delta )$ satisfies
    (5) $$ \begin{align} \frac12\frac{|x^my^n|^d}{1+j|x^my^n|^d}\le |x_j^my_j^n|^d\le \frac 2{\cos\theta}\frac{|x^my^n|^d}{1+j|x^my^n|^d}\ \end{align} $$
    for all $j\ge 0$ . Moreover, any orbit of F that converges to 0 eventually lies in $U_k(\varepsilon ,\theta )$ for some $k\in \{0,\ldots , d-1\}$ .
  2. (2) There exists $\tilde \delta \in (0,1]$ such that $\widetilde D_k(\varepsilon ,\theta ,\tilde \delta )$ is contained in the basin of attraction of $U_k(\varepsilon ,\theta )$ for all $k\in \{0,\ldots , d-1\}$ and

    $$ \begin{align*}F^j(\widetilde D_k(\varepsilon,\theta,\tilde\delta))\subset \widetilde D_k(\varepsilon,\theta,\delta)\end{align*} $$
    for all $j\ge 0$ .

Proof. From the expression for F, we can write $x_1=x(1-\zeta )$ , where $\zeta =(-a+\sigma ) (x^my^n)^d$ and $\sigma =O(x,y)$ . By the choice of $\theta _0$ , we can take $\delta _0>0$ such that for all $(x,y)$ with $|x|<\delta _0$ and $|y|<\delta _0$ we have

$$ \begin{align*}\delta_0<|\!-a+\sigma|< 1/\delta_0, \quad \lvert\arg(-a+\sigma)\rvert+2\theta_0<\pi/2.\end{align*} $$

Set $\eta =\frac 1{2}\delta _0 {\cos (\pi /2-\theta _0)}>0$ and $\varepsilon _0=\delta _0\cos (\pi /2-\theta _0)$ . Let us show that if $\varepsilon \le \varepsilon _0$ and $k\in \{0,\ldots , d-1\}$ then for all $(x,y)\in D_k(\varepsilon ,\theta ,\delta _0)$ we have

(6) $$ \begin{align} |x_1|\le |x|(1-\eta|x^my^n|^d)\le |x|. \end{align} $$

Consider $(x,y)\in D_k(\varepsilon ,\theta ,\delta _0)$ with $\varepsilon \le \varepsilon _0$ . We have that

$$ \begin{align*}|\zeta|<|x^my^n|^d/\delta_0<\varepsilon/\delta_0< 1\end{align*} $$

and

$$ \begin{align*}\lvert\arg\zeta\rvert\le\lvert\arg (-a+\sigma)\rvert+\lvert\arg(x^my^n)^d\rvert<(\pi/2 -2\theta_0 ) +\theta< \pi/2-\theta_0.\end{align*} $$

Thus $|\zeta |<\varepsilon /\delta _0\le \cos (\pi /2 -\theta _0)< \cos (\arg \zeta )$ and so $|\zeta |^2<\operatorname {\mathrm {Re}}\zeta $ . Then

$$ \begin{align*}|1-\zeta|^2=1-2\operatorname{\mathrm{Re}}\zeta+|\zeta|^2<1-\operatorname{\mathrm{Re}}\zeta<1-\operatorname{\mathrm{Re}}\zeta +\tfrac 14 (\operatorname{\mathrm{Re}}\zeta)^2=(1-\tfrac 12 \operatorname{\mathrm{Re}}\zeta)^2\end{align*} $$

and therefore $|1-\zeta |<|1-\tfrac 12 \operatorname {\mathrm {Re}}\zeta |=1-\frac 12 \operatorname {\mathrm {Re}}\zeta $ , so

$$ \begin{align*}|1-\zeta|<1-\tfrac 12 \operatorname{\mathrm{Re}} \zeta= 1-\tfrac 12\cos (\arg\zeta) |\zeta|<1-\tfrac 12\cos (\pi/2 -\theta_0)|\zeta|\end{align*} $$

and hence $|1-\zeta |<1-\frac {1} {2 }\delta _0\cos (\pi /2 -\theta _0) |x^my^n|^d,$ which proves (6). Proceeding in the same way, up to reducing $\delta _0$ if necessary, we get that for all $(x,y)\in D_k(\varepsilon ,\theta ,\delta _0)$ with $\varepsilon \le \varepsilon _0$ ,

(7) $$ \begin{align} |y_1|\le |y|(1-\eta|x^my^n|^d)\le |y|. \end{align} $$

From the expression for F, if $x,y\in \mathbb C^*$ are small enough, we have that

$$ \begin{align*} \dfrac{|y_1|}{|x_1^my_1^n|^\gamma}&=\dfrac{|y||1+b(x^my^n)^d(1+O(x,y))|}{|x^my^n|^\gamma|1-(\gamma/d) (x^my^n)^d(1+O(x,y))|}\\ &=\dfrac{|y|}{|x^my^n|^\gamma}\bigg|1+(x^my^n)^d\bigg[b+\frac{\gamma}{d}+O(x,y)\bigg]\bigg|, \end{align*} $$

so we can write

$$ \begin{align*} \dfrac{|y_1|}{|x_1^my_1^n|^\gamma} =\dfrac{|y|}{|x^my^n|^\gamma}|1-\zeta_1|, \end{align*} $$

where $\zeta _1=(-b-\gamma /d+\tau )(x^my^n)^d$ with $\tau =O(x,y)$ . By the choice of $\theta _0$ , reducing $\delta _0$ if necessary, we get that for all $(x,y)$ with $|x|<\delta _0$ and $|y|<\delta _0$ we have

$$ \begin{align*} \delta_0<|-b-\gamma/d+\tau|<1/\delta_0,\quad \lvert\arg (-b-\gamma/d+\tau)\rvert+2\theta_0<\pi/2. \end{align*} $$

Let us show that if $\varepsilon \le \varepsilon _0$ and $k\in \{0,\ldots , d-1\}$ , then for all $(x,y)\in D_k(\varepsilon ,\theta ,\delta _0)$ we have

(8) $$ \begin{align} \dfrac{|y_1|}{|x_1^my_1^n|^\gamma}\le \dfrac{|y|}{|x^my^n|^\gamma}(1-\eta|x^my^n|^d)\le \dfrac{|y|}{|x^my^n|^\gamma}. \end{align} $$

Consider $(x,y)\in D_k(\varepsilon ,\theta ,\delta _0)$ with $\varepsilon \le \varepsilon _0$ . We have that

$$ \begin{align*}|\zeta_1|<|x^my^n|^d/\delta_0<\varepsilon/\delta_0\le 1 \end{align*} $$

and

$$ \begin{align*}\lvert\arg\zeta_1\rvert\le\lvert\arg (-b-\gamma/d+\tau)\rvert+\lvert\arg(x^my^n)^d\rvert<(\pi/2-2\theta_0 ) +\theta< \pi/2-\theta_0,\end{align*} $$

so proceeding exactly as above, we obtain that $|1-\zeta _1|<1-\eta |x^my^n|^d$ which proves (8). If $n\ge 1$ , analogously we get, up to reducing $\delta _0$ if necessary, that for $(x,y)\in D_k(\varepsilon ,\theta ,\delta _0)$ with $\varepsilon \le \varepsilon _0$ we have

(9) $$ \begin{align} \dfrac{|x_1|}{|x_1^my_1^n|^\gamma}\le \dfrac{|x|}{|x^my^n|^\gamma}(1-\eta|x^my^n|^d)\le \dfrac{|x|}{|x^my^n|^\gamma}. \end{align} $$

Notice that, because of equation (4), F satisfies the hypotheses of Proposition 2.2. Set

$$ \begin{align*}\delta_\theta=\min\{1, \delta_0,\delta_F(\theta)\} \quad \text{and}\quad \varepsilon_\theta=\min\{1, \varepsilon_0,\varepsilon_F(\theta), \delta_0^{d/\gamma}, \delta_F(\theta)^{d/\gamma}\},\end{align*} $$

where $\varepsilon _F(\theta )$ and $\delta _F(\theta )$ are the constants given by Proposition 2.2. Now, consider ${\varepsilon \le \varepsilon _\theta }$ , $\delta \le \delta _\theta $ and $k\in \{0,\ldots , d-1\}$ and let us show that $D_k(\varepsilon , \theta ,\delta )$ and $U_k(\varepsilon ,\theta )$ are invariant by F and that estimates (5) hold. First, take $(x,y)\in D_k(\varepsilon ,\theta ,\delta )$ . Then, it follows from (6) and (7) that $|x_1|\le |x|<\delta $ and $|y_1|\le |y|<\delta $ . Moreover, since $\varepsilon \le \varepsilon _F(\theta )$ , $\delta \le \delta _F(\theta )$ and $x^my^n\in S_k(d,\varepsilon ,\theta )$ , we have by assertion 1 of Proposition 2.2 that $x^m_1y_1^n\in S_k(d,\varepsilon , \theta )$ . Therefore $(x_1,y_1)\in D_k(\varepsilon ,\theta ,\delta )$ and so $D_k(\varepsilon ,\theta ,\delta )$ is invariant by F. Now, take $(x,y)\in U_k(\varepsilon ,\theta )$ . Then $|y|<|x^my^n|^\gamma <\varepsilon ^{\gamma /d}\le \delta _0$ and in the same way $|x|<\delta _0$ , so $(x,y)\in D_k(\varepsilon ,\theta ,\delta _0)$ . It follows from (8) and (9) that $|y_1|<|x_1^my_1^n|^\gamma $ and analogously $|x_1|<|x_1^my_1^n|^\gamma $ . Moreover, since $x^my^n\in S_k(d,\varepsilon ,\theta )$ , $\varepsilon \le \varepsilon _F(\theta )$ , $|y|<\varepsilon ^{\gamma /d}<\delta _F(\theta )$ and $|x|<\varepsilon ^{\gamma /d}<\delta _F(\theta )$ , we have by assertion 1 of Proposition 2.2 that $x_1^my_1^n\in S_k(d,\varepsilon ,\theta )$ , which proves that $U_k(\varepsilon ,\theta )$ is invariant by F. Moreover, if $(x,y)\in U_k(\varepsilon , \delta )\cup D_k(\varepsilon ,\delta ,\theta )$ we have estimates (5) directly from assertion 1 of Proposition 2.2.

Let us prove now that $F^j\to 0$ uniformly on $U_k(\varepsilon ,\theta )$ . Take $(x,y)\in U_k(\varepsilon ,\theta )$ . Since $(x_j,y_j)\in U_k(\varepsilon ,\theta )$ for all $j\ge 0$ , by (5) we have

$$ \begin{align*} |x_j^my_j^n|^d\le \frac 2{\cos\theta}\frac{|x^my^n|^d}{1+j|x^my^n|^d}\le \frac 2{\cos\theta} \frac1j \end{align*} $$

for all $j\ge 0$ , which shows that $x_j^my_j^n\to 0$ uniformly on $U_k(\varepsilon ,\theta )$ . Then, since ${|y_j|<|x_j^my_j^n|^\gamma }$ and $|x_j|<|x_j^my_j^n|^\gamma $ , we have that $x_j\to 0$ and $y_j\to 0$ uniformly on $U_k(\varepsilon ,\theta )$ .

Consider now an orbit $(x_j,y_j)$ converging to 0 and let us show that it eventually lies in $U_k=U_k(\varepsilon ,\theta )$ for some k. Let $\tilde \delta _F(\theta )$ be the constant given by Proposition 2.2. Since $(x_j,y_j)\to 0$ , there exists $j_0\ge 0$ such that $|x_j|<\tilde \delta _F(\theta )$ and $|y_j|<\tilde \delta _F(\theta )$ for all $j\ge j_0$ . By assertion 3 of Proposition 2.2, up to increasing $j_0$ , we have that $x_j^my_j^n\in S_k(d,\varepsilon ,\theta )$ for all $j\ge j_0$ and for some $k\in \{0,\ldots , d-1\}$ so, increasing $j_0$ again if necessary, we get that $(x_j,y_j)\in D_k(\varepsilon ,\theta ,\delta _0)$ for all $j\ge j_0$ . Then, by iterated applications of (8),

$$ \begin{align*} \dfrac{|y_j|}{|x_j^my_j^n|^\gamma}&\le \dfrac{|y_{j_0}|}{|x_{j_0}^my_{j_0}^n|^\gamma}\prod\limits_{l=j_0}^{j-1} (1-\eta|x_l^my_l^n|^d) \end{align*} $$

for all $j\ge j_0$ . The estimate

$$ \begin{align*}|x_j^my_j^n|^d\ge \frac 12 \frac{|x^my^n|^d}{1+j|x^my^n|^d}\end{align*} $$

from (5) implies that $\prod _{l=j_0}^{j-1}(1-\eta |x_l^my_l^n|^d)$ tends to 0 as $j\to +\infty $ and so does $|y_j|/|x_j^my_j^n|^\gamma $ . Hence $|y_j|\le |x_j^my_j^n|^\gamma $ for j big enough and, if $n\ge 1$ , analogously $|x_j|\le |x_j^my_j^n|^\gamma $ for j big enough, which proves that eventually $(x_j,y_j)\in U_k$ . Then, to prove that $D_k(\varepsilon ,\theta ,\delta )$ is contained in the basin of attraction of $U_k(\varepsilon , \theta )$ it suffices to show that the orbit of any point in $D_k(\varepsilon ,\theta ,\delta )$ converges to 0. Take $(x,y)\in D_k(\varepsilon ,\theta ,\delta )$ . Since $(x_j,y_j)\in D_k(\varepsilon ,\theta ,\delta )$ for all $j\ge 0$ , we can apply (6) and (7) for each $(x_j,y_j)$ and we have

$$ \begin{align*}|x_j|\le |x_0|\prod\limits_{l=0}^{j-1}(1-\eta|x_l^my_l^n|^d),\quad |y_j|\le |y_0|\prod\limits_{l=0}^{j-1}(1-\eta|x_l^my_l^n|^d).\end{align*} $$

We have as before, by estimates (5), that the product above tends to zero and so do $x_j$ and $y_j$ . This concludes the proof of assertion 1.

Let us prove assertion 2. From the expression for F we have in a neighborhood of $0\in \mathbb C^2$ that $|x_1|\le |x|(1+\nu |x^my^n|^d)$ and $|y_1|\le |y|(1+\nu |x^my^n|^d)$ for some $\nu>0$ . Let $\tilde \delta =\min \{1,\tilde \mu \}$ , where $\tilde \mu $ is given by assertion 2 of Proposition 2.2 for $\mu =\delta $ . If $x^my^n\in \widetilde D_k(\varepsilon ,\theta ,\tilde \delta )$ , by Proposition 2.2 there exists $j_0\ge 0$ such that $|x_j|<\delta $ , $|y_j|<\delta $ and $ x^m_{j}y^n_{j}\in \widetilde S_k(d,\varepsilon ,\theta )$ for all $j\le j_0$ and $x^m_{j_0}y^n_{j_0}\in S_k(d,\varepsilon ,\theta )$ . In particular, $(x_{j_0},y_{j_0})\in D_k(\varepsilon ,\theta ,\delta )$ so by assertion 1 $(x_j,y_j)\in D_k(\varepsilon ,\theta , \delta )$ for all $j\ge j_0$ so it eventually lies in $U_k(\varepsilon ,\theta )$ . Moreover, since $|x_j|<\delta $ , $|y_j|<\delta $ and $ x^m_{j}y^n_{j}\in \widetilde S_k(d,\varepsilon ,\theta )$ for all $j\ge 0$ , we have that $F^j(\widetilde D_k(\varepsilon ,\theta ,\tilde \delta ))\subset \widetilde D_k(\varepsilon ,\theta ,\delta )$ for all $ j\ge 0.$ This concludes the proof of the proposition.

4 Existence of invariant functions

In this section we show the existence of invariant functions for F on the domains ${U_k=U_k(\varepsilon ,\theta )}$ given by Proposition 3.3 (up to reducing $\varepsilon $ ).

Since F is close to the time-one flow of the vector field

$$ \begin{align*}x^My^N\bigg( ax\dfrac{\partial}{\partial x}+ by\dfrac{\partial}{\partial y}\bigg)\end{align*} $$

and the vector field $ax({\partial }/{\partial x})+ by({\partial }/{\partial y})$ has the Liouvillian first integrals $x^{\eta b}y^{-\eta a}$ , ${\eta \in \mathbb {C}^*}$ , our aim is to find an invariant function close to one of these first integrals, for which we start by defining a suitable branch $g(x,y)$ of $x^{d b}y^{-d a}$ on $U_k$ , where $0\le k\le d-1$ . From now on, if $z\in \mathbb {C}\backslash [-\infty ,0]$ and $\unicode{x3bb} \in \mathbb {C}\backslash \mathbb {Z}$ , we denote $z^{\unicode{x3bb} }=e^{\unicode{x3bb} \log z},$ where $\log $ is the main branch of the logarithm. Note that, since m and n are coprime if $n\ge 1$ , there exist $p,q\in \mathbb {N}$ such that $qm-pn=1$ ; if $n=0$ , we set $p=0$ , $q=1$ . Denote $\unicode{x3bb} =d(ap+bq)$ and define $g:U_k\to \mathbb C$ as

(10) $$ \begin{align} g(x,y)=x^py^{q}(x^my^n)^{\unicode{x3bb}}, \end{align} $$

which is well defined since $x^my^n$ belongs to $\mathbb {C}\backslash [-\infty ,0]$ for all $(x,y)\in U_k$ . We point out that if $n\ge 1$ , since $U_k$ is disjoint from $\{xy=0\}$ , the function g is non-vanishing whereas if $n=0$ , since $g(x,y)=yx^{Mb}$ and $U_k$ intersects $\{y=0\}$ , the function g has $U_k\cap \{y=0\}$ as zero set.

If x and y belong to a small sector bisected by $\mathbb R^+$ we have, in view of the identity $dma+dnb=-1$ , that

$$ \begin{align*}g(x,y)=x^py^{q}(x^my^n)^{\unicode{x3bb}}= x^{p+m\unicode{x3bb}}y^{q+n\unicode{x3bb}}=x^{d b}y^{-d a},\end{align*} $$

so g is a branch of $ x^{d b}y^{-d a}$ on $U_k$ .

Proposition 4.1. Let F be a biholomorphism of the form (B) with $N\ge 0$ satisfying condition (B ) and fix $\theta _0$ as in §3. Consider $\theta \in (0,\theta _0)$ and let $\varepsilon _\theta $ be the constant given by Proposition 3.3. Then there exist $\tilde \varepsilon _\theta \in (0,\varepsilon _\theta ]$ and functions $\psi _k\in \mathcal {O}(U_k)$ for all $k\in \{0,\ldots , d-1\}$ , where $U_k=U_k(\tilde \varepsilon _\theta ,\theta )$ , such that $\psi _k\circ F=\psi _k$ . Moreover, $\psi _k=ug$ , where g is the function defined above and $u\in \mathcal {O}(U_k)$ satisfies $ |u(x,y)-1|<1/2$ for all $(x,y)\in U_k$ ; in particular, $\psi _k$ is non-vanishing if $n\ge 1$ and has $U_k\cap \{y=0\}$ as zero set if $n=0$ .

Proof. Take $\tilde \varepsilon _\theta \le \varepsilon _\theta $ . Note that the domains $U_k=U_k(\tilde \varepsilon _\theta ,\theta )$ for $k\in \{0,\ldots , d-1\}$ satisfy the conclusion of Proposition 3.3.

We define $\psi _k$ as

$$ \begin{align*}\psi_k(x,y)=\lim\limits_{j\to\infty}g(x_j,y_j), \ \ (x,y)\in U_k,\end{align*} $$

where $(x_j,y_j)=F^j(x,y)$ . It is clear that this function, if well defined, will be invariant by F. Let us show that it is well defined and holomorphic. Using the expression for F and equation (4), we have that

$$ \begin{align*}\dfrac{x_1^py_1^q}{x^py^q}=1+(x^my^n)^d[ap+bq+O(x,y)]\end{align*} $$

and

for all $(x,y)$ , so

(11) $$ \begin{align} \frac{g(x_1,y_1)}{g(x,y)}=1+\ell(x,y), \quad \text{with } \ell(x,y)=(x^my^n)^dO(x,y). \end{align} $$

Since $(x_j,y_j)\in U_k$ for all j, we have that $|y_j| \le |x_j^my_j^n|^\gamma $ and $|x_j|\le |x_j^my_j^n|^{\gamma }$ so

$$ \begin{align*}|\ell(x_j,y_j)|\le K|x_j^my_j^n|^{d+\gamma}\end{align*} $$

for some $K>0$ . Therefore, by estimates (5), the product $\prod _{j\ge 0}({g(x_{j+1},y_{j+1})}/g(x_j, y_j))$ converges uniformly for $(x,y)\in {U_k}$ and defines a holomorphic function $u\in \mathcal {O}({U_k})$ . Then $g(x_j,y_j)\to u(x,y)g(x,y)$ uniformly on $U_k$ , so $\psi _k$ is well defined and holomorphic in ${U_k}$ , and we have $\psi _k=ug$ . Note that the function u is arbitrarily close to $1$ if we suppose $|x^my^n|$ to be small enough: if $n\in \mathbb {N}$ is large enough the finite product $\prod _{0\le j\le n}({g(x_{j+1},y_{j+1})}/{g(x_j,y_j)})$ is arbitrarily uniformly close to u and we have from (11) that this finite product is arbitrarily uniformly close to 1 if $|x^my^n|$ is small enough. Therefore, since $|x^my^n|<\tilde \varepsilon _\theta ^{1/d}$ in $U_k$ , we get that $|u(x,y)-1|<1/2$ for all $(x,y)\in U_k$ , provided $\tilde \varepsilon _\theta $ is small enough.

5 Approximate Fatou coordinates

In this section we find a change of coordinates $\phi _k$ on each of the domains $U_k(\varepsilon ,\theta )$ provided by Propositions 3.3 and 4.1 (up to reducing $\varepsilon $ and $\theta $ ) which gives a first approximation of Fatou coordinates (that is, conjugations with $(z,w)\mapsto (z+1,w)$ ).

Definition 5.1. Given $\theta \in (0,\pi /2)$ and $\varepsilon \in (0,1]$ such that Proposition 4.1 holds in ${U_k=U_k(\varepsilon ,\theta )}$ , we consider the map $\phi _k\colon U_k\to \mathbb C^2$ given by

$$ \begin{align*}\phi_k(x,y)=\bigg(\frac{1}{(x^my^n)^d},\psi_k(x,y)\bigg),\end{align*} $$

where $\psi _k\in \mathcal {O}(U_k)$ is the invariant function for F given by Proposition 4.1. Note that $\phi _k(U_k)\subset \mathbb {C}\times \mathbb {C}^*$ if $n\ge 1$ . We also define the set $V=V(\varepsilon ,\theta ,r)$ , for $0<r<1$ , as

$$ \begin{align*}V=\{(z,w)\in\mathbb C^2:|z|>\varepsilon^{-1}, \ \lvert\arg z\rvert<\theta, \ |w|<r|z|^{-\mathrm{Re}\ b/m-\gamma/(dm)}\}\end{align*} $$

if $n=0$ and

$$ \begin{align*}V&=\{(z,w)\in\mathbb C^2:|z|>\varepsilon^{-1}, \ \lvert\arg z\rvert<\theta, \\&\ \ \quad r^{-1}|z|^{\mathrm{Re}\ a/n+\gamma/(dn)} <|w|<r|z|^{-\mathrm{Re}\ b/m-\gamma/(dm)}\}\end{align*} $$

if $n\ge 1$ . Since $\operatorname {\mathrm {Re}} a+\gamma /d<0<-\operatorname {\mathrm {Re}} b-\gamma /d$ , V is non-empty. Notice also that V is homeomorphic to $\mathbb C^2$ if $n=0$ and to $\mathbb C\times \mathbb C^*$ if $n\ge 1$ .

Proposition 5.2. Let F be a biholomorphism of the form (B) with $N\ge 0$ satisfying condition (B ) and fix $\theta _0$ as in §3. There exist $\theta _1\in (0,\theta _0)$ and $\varepsilon _1\in (0,\tilde \varepsilon _{\theta _1}]$ , where $\tilde \varepsilon _{\theta _1}$ is the constant given by Proposition 4.1, such that if $\varepsilon <\varepsilon _1$ , $\theta <\theta _1$ and r is small enough and we denote $V=V(\varepsilon ,\theta ,r)$ then the following properties hold for each $k\in \{0,\ldots , d-1\}$ .

  1. (1) $V\subset \phi _k(U_k)$ and $\phi _k:\phi _k^{-1}(V)\to V$ is a biholomorphism.

  2. (2) $U_k(\varepsilon ,\theta )$ is in the basin of attraction of $ \phi _k^{-1}(V)$ .

  3. (3) The map $\tilde {F}= \phi _k\circ F\circ \phi _k^{-1}$ maps V into V and has the form $\tilde {F}(z,w)=(z+1+h(z,w),w)$ with

    $$ \begin{align*}|z+1+h(z,w)|\kern1.4pt{>}\kern1.4pt |z|+\frac 12, \!\quad\! |h(z,w)|\kern1.4pt{<}\kern1.4pt K|z|^{-\gamma/d} \!\quad\! \text{and} \!\quad\! \bigg|\frac{\partial h}{\partial z} (z,w)\bigg|\kern1.2pt{<}\kern1.2pt K'|z|^{-1-\gamma/d}\end{align*} $$
    for some $K, K'>0$ and for all $(x,y)\in V$ .

Proof. Consider $\theta _1\in (0,\theta _0)$ with $\cos \theta _1>2/3$ and let $\varepsilon _1=\tilde \varepsilon _{\theta _1}$ be the constant given by Proposition 4.1. From equation (4) we have that

$$ \begin{align*} \frac{1}{(x_1^my_1^n)^d}=\frac{1}{(x^my^n)^d}+1+ O(x,y). \end{align*} $$

Thus, since $|x|\le |x^my^n|^\gamma $ and $|y|\le |x^my^n|^{\gamma }$ for all $(x,y)\in U_k(\varepsilon _1,\theta _1)$ , there exists ${K>0}$ such that

(12) $$ \begin{align} \bigg|\frac{1}{(x_1^my_1^n)^d}-\frac{1}{(x^my^n)^d}-1\bigg|<K|x^my^n|^{\gamma}<K\varepsilon_1^{\gamma/d} \end{align} $$

for all $(x,y)\in U_k(\varepsilon _1,\theta _1)$ and all $k\in \{0,\ldots , d-1\}$ . Then, up to reducing $\varepsilon _1$ if necessary, we can assume that

(13) $$ \begin{align} \bigg|\frac{1}{(x_1^my_1^n)^d}-\frac{1}{(x^my^n)^d}-1\bigg|<\frac 16 \end{align} $$

for all $(x,y)\in U_k(\varepsilon _1,\theta _1)$ and all $k\in \{1,\ldots , d-1\}$ . We now fix $\theta <\theta _1$ and $\varepsilon <\varepsilon _1$ and we denote $U_k=U_k(\varepsilon ,\theta )$ .

Let us prove assertion 1; without loss of generality we assume that $k=0$ . Let g be the function defined by (10). Using the fact that $qm-pn=1$ and $adm+bdn=-1$ , a straightforward computation shows that the map $\varphi :U_0\to \mathbb C^2$ given by

$$ \begin{align*}\varphi(x,y)=\bigg(\frac{1}{(x^my^n)^d},g(x,y)\bigg)\end{align*} $$

is injective and its inverse is given by

$$ \begin{align*}\varphi^{-1}(z,w)=(z^aw^{-n},z^bw^{m}).\end{align*} $$

Consider a point $(z_0,w_0)\in V$ , and set

$$ \begin{align*}A_{z_0}=\{(x,y)\in\mathbb C^2: x^my^n=z_0^{-1/d}, \ |\iota x|<|z_0|^{-\gamma/d}, \ |y|<|z_0|^{-\gamma/d}\}.\end{align*} $$

Notice that $A_{z_0}\subset U_0$ and $\overline {A_{z_0}}$ is a Riemann surface with boundary. If we set $x=z_0^aw_0^{-n}$ and $y=z_0^bw_0^{m}$ , then $x^my^n=z_0^{-1/d}$ and $g(x,y)=w_0$ . Moreover, $|y|=|z_0^bw_0^m|<r^m|z_0^b||z_0|^{-\mathrm{Re}\ b-\gamma/d}\le r^me^{|b|\theta }|z_0|^{-\gamma /d}$ and, if $n\ge 1$ , analogously $|x|=|z_0^aw_0^{-n}|<r^ne^{|a|\theta }|z_0|^{-\gamma /d}$ , so $(x,y)\in A_{z_0}$ if r is small enough. Therefore, since $\varphi $ is injective, $g|_{A_{z_0}}$ assumes the value $w_0$ once. If we show that

(14) $$ \begin{align} |\psi_0(x,y)-g(x,y)|<|g(x,y)-w_0|\quad \textrm{whenever} \ (x,y)\in\partial A_{z_0} \end{align} $$

then, by Rouché’s theorem, the function $\psi _0|_{A_{z_0}}$ will also assume the value $w_0$ exactly once, showing that $V\subset \phi _0(U_0)$ and that $\phi _0$ is injective in $\phi _0^{-1}(V)$ . Let us prove that inequality (14) holds. If $n\ge 1$ , the boundary $\partial A_{z_0}$ of $A_{z_0}$ is composed by two connected components,

$$ \begin{align*} \partial_1 A_{z_0}&=\{(x,y)\in\mathbb C^2:x^my^n=z_0^{-1/d}, \ |x|=|z_0|^{-\gamma/d}\},\\ \partial_2 A_{z_0}&=\{(x,y)\in\mathbb C^2:x^my^n=z_0^{-1/d}, \ |y|=|z_0|^{-\gamma/d}\}, \end{align*} $$

whereas if $n=0$ we have $\partial A_{z_0}=\partial _2 A_{z_0}$ . Consider a point $(x,y)\in \partial A_{z_0}\subset U_0$ . Since $x^my^n=z_0^{-1/d}$ , it follows from the computation of $\varphi ^{-1}$ that

$$ \begin{align*}g(x,y)^n= x^{-1}z_0^a\quad\text{and}\quad g(x,y)^m=yz_0^{-b}.\end{align*} $$

We suppose first that $(x,y)\in \partial _2 A_{z_0}$ . Then

$$ \begin{align*}|g(x,y)|= |y|^{{1}/{m}}|z_0^{-b/m}| = |z_0|^{-\gamma/(dm)}|z_0^{-b/m}| \ge e^{-{|b|\theta}/m}|z_0|^{-\operatorname{\mathrm{Re}} b/m-\gamma/(dm)} \end{align*} $$

so $|g(x,y)|>e^{-{|b|\theta }/m}r^{-1} |w_0|$ . Then

$$ \begin{align*}|g(x,y)-w_0|\ge |g(x,y)|-|w_0|>(1-e^{{|b|\theta}/m}r)|g(x,y)|\ge \tfrac{1}{2}|g(x,y)|\end{align*} $$

if r is small enough. This relation, together with the fact that

$$ \begin{align*}|\psi_0(x,y)-g(x,y)|<\tfrac12|g(x,y)|\end{align*} $$

for all $(x,y)\in U_0$ , as was shown in Proposition 4.1, implies (14). Analogously, if $(x,y)\in \partial _1 A_{z_0}$ (so $n\ge 1$ ) then

$$ \begin{align*}|g(x,y)|=|x|^{-{1}/{n}}|z_0^{a/n}|= |z_0|^{\gamma/(dn)}|z_0^{{a}/{n}}| \le e^{{|a|\theta}/{n}}|z_0|^{{\operatorname{\mathrm{Re}} a}/{n}+{\gamma}/{(dn)}}\end{align*} $$

so $|g(x,y)|< e^{{|a|\theta }/{n}}r|w_0|$ and hence

$$ \begin{align*}|g(x,y)-w_0|\ge|w_0|- |g(x,y)|>(e^{-{|a|\theta}/{n}}r^{-1}-1)|g(x,y)| \ge \tfrac{1}{2}|g(x,y)|\end{align*} $$

if r is sufficiently small, which again implies (14) and assertion 1 is proved.

Now take $(x,y)\in U_k$ . If we set $(z_j,w_j)=\phi _k(x_j,y_j)$ , then clearly $|z_j|>\varepsilon ^{-1}$ and $\lvert \arg z_j\rvert <\theta $ ; moreover, since $\psi _k$ is invariant by F we have that $w_j$ is constant for all j while $z_j\to +\infty $ because of (5), so for j large enough we get $|w_j|<r|z_j|^{{-\operatorname {\mathrm {Re}} b}/{m}-{\gamma }/{(dm)}}$ and if $n\ge 1$ , since $w_j\neq 0$ , $|w_j|>r^{-1}|z_j|^{{\operatorname {\mathrm {Re}} a}/{n}+\gamma /{(dn)}}$ , so $(z_j,w_j)\in V$ and assertion 2 is proved.

Let us prove assertion 3. For $r_1\in (0,1)$ , consider the set $V_1=V(\varepsilon _1,\theta _1,r_1)$ . Proceeding exactly as in the proof of assertion 1, we can choose $r_1\in (0,1)$ such that $V_1 \subset \phi _k(U_k(\varepsilon _1,\theta _1))$ and $\phi _k:\phi _k^{-1}(V_1)\to V_1$ is a biholomorphism for every k. Then $\phi _k^{-1}$ is well defined on $V_1$ and takes values in $U_k(\varepsilon _1,\theta _1)$ , so $\tilde {F}= \phi _k\circ F\circ \phi _k^{-1}$ is well defined on $V_1$ . Since $\psi _k$ is invariant by F, we can express $\tilde {F}(z,w)=(f(z,w),w)$ . Then, if we write $\phi _k^{-1}(z,w)=(x,y)\in U_k(\varepsilon _1,\theta _1)$ and $F(x,y)=(x_1,y_1)$ , we have from equation (4) that

$$ \begin{align*}f(z,w)=\frac{1}{(x_1^my_1^n)^d}=\frac{1}{(x^my^n)^d}+1+\tilde{h}(x,y),\end{align*} $$

where $\tilde {h}(x,y)=O(x,y)$ . That is,

$$ \begin{align*}f(z,w)=z+1+{h}(z,w),\end{align*} $$

where by (12) and (13)

$$ \begin{align*}|h(z,w)|<K|z|^{-\gamma/d} \quad \text{and} \quad |h(z,w)|<\tfrac 16\end{align*} $$

for all $(z,w)\in V_1$ . Thus, since $\cos \theta _1>2/3$ , we have

$$ \begin{align*} |f(z,w)|^2&=|z+1+h(z,w)|^2= |z|^2+2\operatorname{\mathrm{Re}} z +2\operatorname{\mathrm{Re}}(z\overline{h}(z,w))+|1+h(z,w)|^2 \\ &\ge |z|^2+2\cos\theta_1|z|-2|z||h(z,w)|+|1+h(z,w)|^2\\ &> |z|^2+4|z|/3-|z|/3+(5/6)^2>(|z|+1/2)^2, \end{align*} $$

so $|f(z,w)|> |z| +1/2$ for all $(z,w)\in V_1$ .

We now fix r satisfying assertion 1 and such that $r<r_1$ . Consider a point $(z,w)\in V$ . Since $(f(z,w),w)\in \phi _k(U_k)$ , it is clear that $|f(z,w)|>\varepsilon ^{-1}$ and $\lvert \arg (f(z,w))\rvert <\theta $ and, since $|w|<r|z|^{{-\operatorname {\mathrm {Re}} b}/{m}-{\gamma }/{(dm)}}$ and $|f(z,w)|>|z|$ , we also have that

$$ \begin{align*}|w|<r|f(z,w)|^{{-\operatorname{\mathrm{Re}} b}/{m}-{\gamma}/{(dm)}}.\end{align*} $$

Analogously, if $n\ge 1$ , since $|w|>r^{-1}|z|^{{\operatorname {\mathrm {Re}} a}/{n}+\gamma /{(dn)}}$ and $|f(z,w)|>|z|$ , we have $|w|>r^{-1}|f(z,w)|^{{\operatorname {\mathrm {Re}} a}/{n}+\gamma /{(dn)}}$ , so $\tilde F(z,w)\in V$ .

To prove the bound for ${\partial h}/{\partial z}$ let us first show that there exists $\rho>0$ such that if $(z_0,w_0)\in V$ and $|z-z_0|<\rho |z_0|$ then $(z,w_0)\in V_1$ . Consider $(z_0,w_0)\in V$ and assume that $|z-z_0|<\rho |z_0|$ . Then

$$ \begin{align*}|z|>(1-\rho )|z_0|>(1-\rho)\varepsilon^{-1},\end{align*} $$

so $|z|>\varepsilon _1^{-1}$ for $\rho $ sufficiently small. Since $|z/z_0-1|<\rho $ , we have $\lvert \arg (z/z_0)\rvert <\arcsin \rho $ , so

$$ \begin{align*}\lvert\arg z\rvert\le \lvert\arg z_0 \rvert+\arcsin\rho<\theta +\arcsin\rho,\end{align*} $$

hence $\lvert \arg z\rvert <\theta _1$ if $\rho $ is small enough. Since $|z_0|<(1-\rho )^{-1}|z|$ , it follows that

$$ \begin{align*}|w_0|<r|z_0|^{{-\operatorname{\mathrm{Re}} b}/{m}-{\gamma}/{(dm)}}<r[(1-\rho)^{-1}|z|]^{{-\operatorname{\mathrm{Re}} b}/{m}-{\gamma}/{(dm)}},\end{align*} $$

so $|w_0|<r_1|z|^{{-\operatorname {\mathrm {Re}} b}/{m}-{\gamma }/{(dm)}}$ if $\rho $ is small enough and, if $n\ge 1$ ,

$$ \begin{align*}|w_0|>r^{-1}|z_0|^{{\operatorname{\mathrm{Re}} a}/{n}+\gamma/{(dn)}}>r^{-1}[(1-\rho)^{-1}|z|]^{{\operatorname{\mathrm{Re}} a}/{n}+\gamma/{(dn)}},\end{align*} $$

so $|w_0|>r_1^{-1}|z|^{{\operatorname {\mathrm {Re}} a}/{n}+\gamma /{(dn)}}$ if $\rho $ is small enough. Hence, $(z,w_0)\in V_1$ if $\rho $ is small enough. Now, take a point $(z_0,w_0)\in V$ . As we have seen, if $D\subset \mathbb {C}$ is the disk of radius $\rho |z_0|$ centered at $z_0$ , then $D\times \{w_0\}$ is contained in $V_1$ , so the function

$$ \begin{align*}h_{w_0}\colon z\in D \mapsto h(z,w_0)\end{align*} $$

is well defined and

$$ \begin{align*}|h_{w_0}(z)|< K|z|^{-\gamma/d}< K(1-\rho)^{-\gamma/d}|z_0|^{-\gamma/d}.\end{align*} $$

Thus, it follows from Cauchy’s inequality that

$$ \begin{align*} \bigg|\frac{\partial h}{\partial z}(z_0,w_0)\bigg|=|(h_{w_0})'(z_0)|\le K(1-\rho)^{-\gamma/d}|z_0|^{-\gamma/d}(\rho|z_0|)^{-1}=K'|z_0|^{-1-\gamma/d}, \end{align*} $$

which finishes the proof of assertion 3.

6 Existence of Fatou coordinates

In this section we construct Fatou coordinates for F on the domains $\phi _k^{-1}(V)\subset U_k$ given by Proposition 5.2.

Proposition 6.1. Let F be a biholomorphism of the form (B) with $N\ge 0$ satisfying condition (B ), let $\varepsilon $ , $\theta $ and r be as in Proposition 5.2 and denote $U_k=U_k(\varepsilon ,\theta )$ and $V=V(\varepsilon ,\theta , r)$ . Then, for each $k\in \{0,\ldots ,d-1\}$ , there exists a biholomorphism $\varphi ^+_k\colon \phi _k^{-1}(V) \to W\subset \mathbb C^2$ , with $W\subset \mathbb C\times \mathbb C^*$ if $n\ge 1$ , conjugating ${F}$ with the map $(z,w)\mapsto (z+1,w)$ and satisfying

(15) $$ \begin{align} \bigcup_{j\in\mathbb N} [W-(j,0)]=\mathbb C^2 \quad \text {if } n=0;\quad \bigcup_{j\in\mathbb N} [W-(j,0)]=\mathbb C\times\mathbb C^* \text { if } n\ge1. \end{align} $$

Proof. By Proposition 5.2, we have that $\phi _k:\phi _k^{-1}(V)\to V$ is a biholomorphism conjugating F with $\widetilde F=\phi _k\circ F\circ \phi _k^{-1}\colon V\to V$ . Thus, it is enough to find a biholomorphism $\Phi \colon V\to W\subset \mathbb C^2$ , with $W\subset \mathbb C\times \mathbb C^*$ if $n\ge 1$ , conjugating $\tilde {F}$ with the map $(z,w)\mapsto (z+1,w)$ . Since $\tilde {F}\colon V\to V$ is written $\tilde {F}(z,w)=(f(z,w),w)$ , each function $f_{w}\colon z\mapsto f(z,w)$ maps the domain $V_w=\{z\in \mathbb C:(z,w)\in V\}$ into itself. Thus, we start considering $w\in \mathbb C$ fixed ( $w\in \mathbb {C}^*$ if $n\ge 1$ ) and we will find, following the ideas in [Reference Milnor13, Lemma 10.10], a map $\beta _w:V_w\to \mathbb C$ conjugating $f_w$ with $z\mapsto z+1$ . We will also show, arguing as in [Reference Milnor13, Lemma 10.11], that

(16) $$ \begin{align}\bigcup_{j\in\mathbb{N}} (\beta_w(V_w)-j)=\mathbb{C}. \end{align} $$

Finally, from the maps $\beta _w$ we will construct a global map $\beta :V\to \mathbb C$ such that ${\beta \circ \tilde F(z,w)=\beta (z,w)+1}$ , so the function $\Phi :V\to \mathbb C^2$ given by $\Phi (z,w)=(\beta (z,w),w)$ is a Fatou coordinate for $\tilde F$ .

From the definition of V we have

$$ \begin{align*}V_w=\{z\in\mathbb{C}\colon |z|>R_w,\; \lvert\arg(z)\rvert<\theta \},\end{align*} $$

where

$$ \begin{align*}R_w=\max \{\varepsilon^{-1},\; (r^{-1}|w|)^{-dm/(d\operatorname{\mathrm{Re}} b+\gamma)}, \; \iota(r|w|)^{dn/(d\operatorname{\mathrm{Re}} a+\gamma)}\}\end{align*} $$

(recall that $\iota =0$ if $n=0$ and $ \iota =1$ if $n\ge 1$ ).

Take a base point $p\in V_w$ . The map $\beta _w$ conjugating $f_w$ with $z\mapsto z+1$ will be constructed as the limit of the functions

$$ \begin{align*}\beta_j(z)= f^j_w(z)-f^{j}_w(p),\quad j\in\mathbb{N}.\end{align*} $$

In order to simplify the proof of the convergence of these functions we assume p to be large enough so that for all $z\in V_w$ the Euclidean segment $[z,p]$ is contained in $V_w$ , which is possible because $\theta <\pi /2$ . Since $|f_w(p)|\ge |p|+1/2$ by Proposition 5.2, the sequence $|f_w^j(p)|$ is increasing, hence the property above also holds for $f_w^j(p)$ . In particular we have, for all $z\in V_w$ and all $j\in \mathbb {N}$ ,

$$ \begin{align*}[f_w^j(z),f_w^j(p)]\subset V_w.\end{align*} $$

Since $f_w(z)=z+1+h(z,w)$ and $|{\partial h}/{\partial z}(z,w)|<K'|z|^{-1-\gamma /d}$ by Proposition 5.2, it follows from the mean value inequality that, if $[z_1,z_2]\subset V_w$ , then

$$ \begin{align*} \bigg|\frac{f_w(z_1)-f_w(z_2)}{z_1-z_2}-1\bigg|= \bigg|\frac{h(z_1,w)-h(z_2,w)}{z_1-z_2}\bigg|\le \max\limits_{z\in[z_1,z_2]}\frac{K'}{|z|^{1+\gamma/d}}. \end{align*} $$

Since the angle between $z_1\mathbb {R}^+$ and $z_2\mathbb {R}^+$ is bounded by $2\theta <\pi $ , there is a constant $\tau>0$ depending only on $\theta $ such that $\min \{|z|:z\in [z_1,z_2]\}\ge \tau \min \{|z_1|,|z_2|\}$ , so

$$ \begin{align*}\bigg|\frac{f_w(z_1)-f_w(z_2)}{z_1-z_2}-1\bigg|\le \frac{K'}{(\tau\min\{|z_1|,|z_2|\} )^{1+\gamma/d}}.\end{align*} $$

In particular, setting $z_1=f_w^j(z)$ and $z_2=f_w^j(p)$ , we obtain

$$ \begin{align*}\bigg|\frac{\beta_{j+1}(z)}{\beta_j(z)}-1\bigg|=\bigg|\frac{f_w^{j+1}(z)-f_w^{j+1}(p)}{f_w^{j}(z)-f_w^{j}(p)}-1\bigg| \le \frac{K'}{(\tau\min\{|f_w^{j}(z)|,|f_w^{j}(p)|\})^{1+\gamma/d}}\end{align*} $$

for all $z\in V_w$ and $j\in \mathbb {N}$ . Therefore, since $|f_w^{j}(z)|\ge j/2$ and $|f_w^{j}(p)|\ge j/2$ , we obtain

$$ \begin{align*}\bigg|\frac{\beta_{j+1}(z)}{\beta_j(z)}-1\bigg|\le K'(2\tau^{-1})^{1+\gamma/d}j^{-1-\gamma/d}\end{align*} $$

for all $z\in V_w$ and $j\in \mathbb {N}$ . This shows that the product $\prod ({\beta _{j+1}(z)}/{\beta _j(z)})$ is uniformly convergent in $V_w$ and therefore $\beta _j$ converges uniformly to a function $\beta _w\in \mathcal {O}(V_w)$ . Let us show that $\beta _w$ is one-to-one and conjugates $f_w$ with the map $z\mapsto z+1$ . Since $f_w(z)=z+1+h(z,w)$ and $|h(z,w)|<K|z|^{-\gamma /d}$ by Proposition 5.2, we have

$$ \begin{align*}| f_w^{j+1}(p)-f_w^j(p) -1|= | h(f_w^j(p),w)|\le \frac{K}{|f_w^j(p)|^{\gamma/d}}\to 0 \quad \textrm{as } j\to\infty.\end{align*} $$

Thus, since $\beta _j(f_w(z))=\beta _{j+1}(z)+f_w^{j+1}(p)-f_w^j(p)$ , we obtain, taking $j\to \infty $ ,

$$ \begin{align*}\beta_w(f_w(z))=\beta_w(z)+1,\quad z\in V_w.\end{align*} $$

Finally, since $\beta _j$ is injective for all j and $\beta _w$ is not constant, we conclude that $\beta _w$ is injective.

Now, as in [Reference Milnor13, Lemma 10.11], we prove that $\beta _w$ satisfies (16). We show first that $\lim \nolimits _{z\to \infty }({\beta _w(z)}/{z})=1$ . Since $\beta _j$ tends uniformly to $\beta _w$ , for some $l\in \mathbb {N}$ we have that $|\beta _w-\beta _l|$ is bounded, whence

$$ \begin{align*}|\beta_w-f_w^l|\le |\beta_w-\beta_l|+|\beta_l-f_w^l|\end{align*} $$

is bounded. Then, since $f_w(z)=z+1+h(z,w)$ and $|h(z,w)|<K|z|^{-\gamma /d}$ ,

$$ \begin{align*}\lim\limits_{z\to \infty}\frac{\beta_w(z)}{z}=\lim\limits_{z\to \infty}\frac{f_w^l(z)}{z} =1.\end{align*} $$

Consider $\zeta \in \mathbb {C}$ . In order to prove (16) we will show that for $j\in \mathbb {N}$ large enough the point $\zeta _j=\zeta +j$ belongs to $\beta _w(V_w)$ . Since $V_w$ is essentially a sector of opening $2\theta $ , if we take a positive number $\rho <\sin \theta $ , it is not difficult to see that, for j large enough, the closed disk $D_j$ of radius $r_j = \rho | \zeta _j|$ centered at $\zeta _j$ is contained in $V_w$ . By Rouché’s theorem, if

$$ \begin{align*}|\beta_w(z)-z|<r_j\end{align*} $$

for all $z\in \partial D_j$ , then $\zeta _j\in \beta _w(D_j)\subset \beta _w(V_w)$ . Since $\beta _w(z)/z\to 1$ when $z\to \infty $ , for z large enough we have $|\beta _w(z)-z|<\tau |z|$ , where $\tau>0$ is taken such that ${\tau }(1+\rho )<\rho $ . Then, if $z\in \partial D_j$ and j is large enough,

$$ \begin{align*}|\beta_w(z)-z|<\tau|z|\le\tau(|\zeta_j|+r_j)=\tau(1+\rho)|\zeta_j|<\rho|\zeta_j|=r_j\end{align*} $$

and (16) follows.

The function $\beta _w$ that we have constructed depends on the choice of the base point $p\in V_w$ , but its derivative does not, as we can check from the definition of $\beta _j$ :

$$ \begin{align*}(\beta_w)'(z)=\lim\limits_{j\to\infty}(f_w^j)'(z).\end{align*} $$

It is easy to see that the same choice of p also works for any $w'$ in a neighborhood of w and the function $\beta _{w'}$ will depend holomorphically on $w'$ . That is, $\beta _{w'}(z)$ is a holomorphic function of $(z,w')$ . Thus, we can find an open covering $\mathbb {C}= \bigcup _{i\in I} W_i$ if $n=0$ or ${\mathbb {C}^*= \bigcup _{i\in I} W_i}$ if $n\ge 1$ and, for each $i\in I$ , a holomorphic function

$$ \begin{align*}\beta_i(z,w),\quad \text{for } z\in V_w,\; w\in W_i\end{align*} $$

such that, for each $w\in W_i $ , the map $z\in V_w\mapsto \beta _i(z,w)\in \mathbb {C}$ is univalent and satisfies $\beta _i(f(z,w),w)=\beta _i(z,w)+1.$ Moreover, from the observation above, the partial derivative ${\partial \beta _i}/{\partial z}$ does not depend on $i\in I$ , that is,

$$ \begin{align*}\frac{\partial \beta_i}{\partial z}(z,w)=\frac{\partial \beta_j}{\partial z}(z,w)\quad \text{for } z\in V_w,\; w\in W_i\cap W_j,\;i,j\in I.\end{align*} $$

Therefore, if $W_i\cap W_j\neq \emptyset $ , there is a function $g_{ij}\in \mathcal {O}(W_i\cap W_j)$ such that

(17) $$ \begin{align} \beta_j(z,w)-\beta_i(z,w)= g_{ij}(w) \quad \text{for } z\in V_w,\; w\in W_i\cap W_j, \end{align} $$

hence $g_{ij}+g_{jk}+g_{ki}=0$ on $W_i\cap W_j\cap W_k$ , for $i,j,k\in I$ . Then, since the first Cousin problem can be solved in $\mathbb C$ and $\mathbb {C}^*$ , there exist functions $g_i\in \mathcal {O}(W_i)$ , $i\in I$ , such that $g_{ij}=g_i-g_j$ on $W_i\cap W_j$ and it follows from (17) that

$$ \begin{align*}\beta_j(z,w)+g_j(w)=\beta_i(z,w) +g_{i}(w) \quad\text{for } z\in V_w,\; w\in W_i\cap W_j.\end{align*} $$

Therefore we can define a global function $\beta \in \mathcal {O}(V)$ by

$$ \begin{align*}\beta(z,w)=\beta_i(z,w)+g_i(w)\quad\text{for } z\in V_w,\;w\in W_i,\end{align*} $$

and we can see that for each $w\in \mathbb {C}^*$ , the map

$$ \begin{align*}z\in V_w\mapsto \beta(z,w)\in\mathbb{C}\end{align*} $$

is univalent and $\beta (f(z,w),w)=\beta (z,w)+1$ for every $(z,w)\in V$ . Now it is easy to check that the holomorphic function

$$ \begin{align*}\Phi(z,w)= (\beta(z,w),w), \quad (z,w)\in V,\end{align*} $$

is univalent and satisfies $\Phi \circ \tilde {F}(z,w)=\Phi (z,w)+(1,0)$ for every $(z,w)\in V$ . To show that $W=\Phi (V)$ satisfies (15), consider a point $(z_0,w_0)\in \mathbb C^2$ , with $(z_0,w_0)\in \mathbb {C}\times \mathbb {C}^*$ if $n\ge 1$ . If $w_0\in W_i$ , we have

$$ \begin{align*}\beta(z,w_0)=\beta_i(z,w_0)+g_i(w_0)=\beta_{w_0}(z)+g_i(w_0)\end{align*} $$

for all $z\in V_{w_0}.$ By (16) there exist $z\in V_{w_0}$ and $j\in \mathbb {N}$ such that $[z_0-g_i(w_0)]+j=\beta _{w_0}(z)$ , thus

$$ \begin{align*}\Phi(z,w_0)=(\beta(z,w_0),w_0)=(z_0+j,w_0)\end{align*} $$

and therefore $(z_0,w_0)\in W-(j,0)$ .

7 The flower theorem

In this section we prove Theorem 1.2. We do this first in the case we have been dealing with in the previous sections: a biholomorphism F of the form (B) with $N\ge 0$ and satisfying condition (B ). At the end of the section we deal with biholomorphisms of the form (A).

Let $\varepsilon ,\theta $ be as in Proposition 5.2 and denote $U_k=U_k(\varepsilon ,\theta )$ . Let $\tilde \delta>0$ be the constant given by assertion 2 of Proposition 3.3 and consider the sets $\widetilde D_k(\varepsilon ,\theta ,\tilde \delta )$ . Since $\gcd (m,n)=1$ , it is easy to see that these sets are connected. For each $k\in \{0,\ldots , d-1\}$ we define

$$ \begin{align*}\Omega^+_k=\bigcup_{j\ge0}F^j(\widetilde D_k(\varepsilon,\theta,\tilde\delta)).\end{align*} $$

This set is connected and invariant by F. Moreover, in view of Proposition 3.3, we have that $\Omega ^+_k$ is in the basin of attraction of $U_k$ and

$$ \begin{align*} \widetilde D_k(\varepsilon,\theta,\tilde\delta)\subset \Omega^+_k\subset\widetilde D_k(\varepsilon,\theta,\delta_\theta). \end{align*} $$

It is easy to see that the diffeomorphism $F^{-1}$ is also of the form (B), with the same pair $(M,N)$ and $(-a,-b)$ instead of $(a,b)$ . Thus, if we work with $F^{-1}$ instead of F, our constructions allow us to find connected open sets $\Omega ^-_0, \ldots ,\Omega ^-_{d-1},$ which play for $F^{-1}$ the role of the sets $\Omega ^+_k$ in the case of F. We can assume that the constructions are done with the same constants $\varepsilon $ , $\delta $ , $\tilde \delta $ , etc. Thus, the sets $\Omega ^-_k$ are defined by

$$ \begin{align*}\Omega^-_k=\bigcup_{j\ge0}F^{-j}(\widetilde D^-_k(\varepsilon,\theta,\tilde\delta)),\end{align*} $$

where

$$ \begin{align*} \widetilde D^-_k(\varepsilon,\theta,\tilde\delta)=\{(x,y)\in\mathbb C^2:x^my^n\in \widetilde S^-_k(d,\varepsilon,\theta), |\iota x|<\tilde\delta,|y|<\tilde\delta\}, \end{align*} $$

in which $\widetilde S^-_k(d,\varepsilon ,\theta )$ is one of the connected components of

$$ \begin{align*} \{z\kern1.4pt{\in}\kern1.4pt\mathbb C:|z^d|\kern1.4pt{<}\kern1.4pt\varepsilon, \lvert\arg(z^d)\kern1.4pt{-}\kern1.4pt\pi\rvert\kern1.4pt{<}\kern1.4pt\theta\}\kern1.4pt{\cup}\kern1.4pt \bigg\{ \bigg|z^d\kern1.4pt{-}\kern1.4pt\dfrac{\varepsilon}2 e^{i(\pi\kern1.4pt{-}\kern1.4pt\theta)}\bigg|\kern1.4pt{<}\kern1.4pt\dfrac{\varepsilon}2\bigg\}\kern1.4pt{\cup}\kern1.4pt \bigg\{\bigg|z^d{-}\kern1.4pt\dfrac{\varepsilon}2 e^{i(\pi+\theta)}\bigg|\kern1.4pt{<}\kern1.4pt\dfrac{\varepsilon}2\bigg\}. \end{align*} $$

In each $\Omega _k^-$ , we have that $F^{-j}\to 0$ and

$$ \begin{align*} \widetilde D^-_k(\varepsilon,\theta,\tilde\delta)\subset \Omega^-_k\subset\widetilde D^-_k(\varepsilon,\theta,\delta_\theta). \end{align*} $$

Since the opening of the sets $\widetilde S_k(d,\varepsilon ,\theta )$ and $\widetilde S_k^-(d,\varepsilon ,\theta )$ is greater than $\pi /d$ , it is clear that the domains $\Omega ^+_0,\ldots ,\Omega ^+_{d-1},\Omega ^-_0,\ldots ,\Omega ^-_{d-1}$ , together with the fixed point set $\{xy^\iota =0\}$ , cover the open set

$$ \begin{align*}\{(x,y)\in\mathbb C^2:|x^my^n|<\varepsilon^{1/d}, |\iota x|<\tilde\delta, |y|<\tilde\delta\},\end{align*} $$

so assertions 1 and 2 of Theorem 1.2 are proved.

For each k, let $\varphi ^+_k\colon \phi _k^{-1}(V)\to W\subset \mathbb C^2$ be the biholomorphism given by Proposition 6.1, which conjugates F with the map $(z,w)\mapsto (z+1,w)$ . It is straightforward to extend $\varphi ^+_k$ as a biholomorphism

$$ \begin{align*}\varphi^+_k\colon \Omega^+_k\to W_k^+\subset\mathbb C^2,\end{align*} $$

with $ W_k^+\subset \mathbb C\times \mathbb C^*$ if $n\ge 1$ , defining, for each $p\in \Omega _k^+$ ,

$$ \begin{align*}\varphi^+_k(p)= \varphi^+_k(F^j(p))-(j,0)\end{align*} $$

for any $j\ge 0$ such that $F^j(p)\in \phi _k^{-1}(V)$ ; this is possible because $\Omega _k^+$ is in the basin of attraction of $U_k$ and, by Proposition 5.2, $U_k$ is in the basin of attraction of $\phi _k^{-1}(V)$ . This shows assertion 3a of Theorem 1.2 for the domains $\Omega ^+_0,\ldots ,\Omega ^+_{d-1}$ ; property 3b follows from (15). We can proceed analogously with the sets $\Omega ^-_0,\ldots ,\Omega ^-_{d-1}$ and this finishes the proof of Theorem 1.2 for F of the form (B) with $N\ge 0$ .

Suppose now that F is of the form (A) satisfying (A ). Up to a linear change of coordinates of the form $(x,y)\mapsto (\alpha x,y)$ , we can assume that $a\in \mathbb R^-$ . In this case, Hakim proved in [Reference Hakim8] that if r is small enough then for any $k\in \{0,\ldots ,M-1\}$ there exists a holomorphic map $u_k:D_{r,k}\to \mathbb C$ , with $|u_k(x)|\le K|x\log x|$ for some $K>0$ and for all $x\in D_{r,k}$ , such that $u_k(F_1(x,u_k(x)))=F_2(x,u_k(x))$ , where $D_{r,k}$ is the component of $\{x\in \mathbb C:|x^M-r|<r\}$ bisected by $e^{2\pi i k/M}\mathbb R^+$ . Moreover, with the small modification of her proof introduced in [Reference López-Hernanz11, Lemma 4.4], we can enlarge the domain of definition of $u_k$ to the set $\widetilde S_k(M,\varepsilon _0,\theta )$ for any $\theta \in (0,\pi /2)$ and for $\varepsilon _0$ small enough. Then, making the sectorial change of coordinates

$$ \begin{align*} (x,y)\in \widetilde S_k(M,\varepsilon_0,\theta)\times \mathbb C\mapsto (x,z^k)=(x,y-u_k(x)), \end{align*} $$

we can write

$$ \begin{align*}F(x,z^k)=(x+x^{M+1}[a+O_1(x,z^k)], z^k+x^Mz^k[b+O_1(x,z^k)]),\end{align*} $$

where we use the notation $O_1(x,z^k)=O(x,x\log x,z^k)$ . The key point to note is that all the constructions we made in the previous sections to obtain the invariant sets $\Omega ^+_k$ for a map F of the form (B) with $N=0$ were performed in $\widetilde S_k(M,\varepsilon ,\theta )\times \mathbb C$ , and all the calculations involved work similarly if we have $O_1(x,y)$ instead of $O(x,y)$ . Then, for some $\varepsilon \le \varepsilon _0$ , the domains $\Omega ^+_k$ can be defined in the same way, but in sectorial coordinates $ (x,z^k)$ depending on k, and the same holds for $\Omega _k^-$ . Assertions 1 and 3 of Theorem 1.2, since referred to a fixed $k\in \{0,\ldots , M-1\}$ , follow exactly as above if we work in the corresponding sectorial coordinates. For the proof of assertion 2, it is enough to show that each $\Omega ^+_k$ contains, in the original coordinates $(x,y)$ , a set of the form

$$ \begin{align*} D_k'=\{(x,y)\in\mathbb C^2:x\in\widetilde S_k(M,\varepsilon',\theta),|y|<\delta'\}\end{align*} $$

for some $\varepsilon ',\delta '>0$ , and the same for each $\Omega _k^-$ . By definition, $\Omega _k^+$ contains the set

$$ \begin{align*}\widetilde D_k=\{(x,z^k)\in\mathbb C^2:x\in \widetilde S _k(M,\varepsilon,\theta),\ |z^k|<\tilde \delta\}.\end{align*} $$

Take $\delta '\le \tilde \delta /2$ and let $\varepsilon '\le \varepsilon $ be such that $|u_k(x)|<\tilde \delta /2$ for all $x\in S _k(M,\varepsilon ',\theta )$ and $k\in \{0,\ldots , d-1\}$ . Take $(x,y)\in D^{\prime }_k$ . In the sectorial coordinates this point is given by $(x,z^k)$ with $z^k=y-u_k(x)$ . Then $|z^k|\le |y|+|u_k(x)|<\delta '+\tilde \delta /2\le \tilde \delta $ , so $ (x,z^k)\in \widetilde D_k$ . This proves that $\Omega ^+_k$ contains $D_k'$ and clearly the analogous property holds for each $\Omega _k^-$ , so Theorem 1.2 is proved.

8 Proof of Theorem 1.3

Suppose first that F is of the form (B). As in §3, up to a linear change of coordinates, we can write F as

$$ \begin{align*}F(x,y)=(x+x^{M+1}y^N[a+O(x,y)], y+x^My^{N+1}[b+O(x,y)]),\end{align*} $$

with $M\ge 1$ , $N\ge 1$ and $aM+bN=-1$ , so the hypothesis of Theorem 1.3 means that either $\operatorname {\mathrm {Re}} a>0$ or $\operatorname {\mathrm {Re}} b>0$ . We assume without loss of generality that $\operatorname {\mathrm {Re}} b>0$ . Also as in §3, put $d=\gcd (M,N)$ and set $m=M/d$ , $n=N/d$ . Since $aM+bN=-1$ , we have that

$$ \begin{align*}x_1^my_1^n=x^my^n-\frac 1d(x^my^n)^{d+1}+(x^my^n)^{d+1}O(x,y).\end{align*} $$

Since $y_1= y(1+[b+O(x,y)](x^my^{n})^d)$ and $\operatorname {\mathrm {Re}} b>0$ , with a similar argument to the one we used for equation (6) we find $\varepsilon ,\delta ,\nu>0$ and $\theta \in (0,\pi /2)$ such that, for each ${k\in \{0,\ldots ,d-1\}}$ , if $x^my^n\in S_k(d,\varepsilon ,\theta )$ with $|x|<\delta $ and $|y|<\delta $ then

(18) $$ \begin{align} |y_1|\ge |y|(1+\nu |x^my^{n}|^d). \end{align} $$

Take $\delta '=\min \{\delta ,\tilde \delta _F(\theta )\}$ , where $\tilde \delta _F(\theta )$ is given by Proposition 2.2, and set

$$ \begin{align*}\mathcal{U}=\{(x,y)\in\mathbb C^2: |x|<\delta', |y|<\delta'\}.\end{align*} $$

Consider a point $(x,y)\in \mathcal {U}$ outside the fixed set $\{xy=0\}$ and let us show that there exists $j\in \mathbb N$ such that $(x_j,y_j)\not \in \mathcal {U}$ . Assume by contradiction that $(x_j,y_j)\in \mathcal {U}$ for all j. Then, by assertion 3 of Proposition 2.2, we find $k\in \{0,\ldots ,d-1\}$ and $j_0\ge 0$ such that ${x_j^my_j^n\in S_k(d,\varepsilon ,\theta )}$ for all $j\ge j_0$ . Thus, by iterated applications of (18) we obtain, for all $j\ge j_0$ , that

$$ \begin{align*}|y_j|\ge |y_{j_0}|\prod_{l=j_0}^{j-1}(1+\nu |x_l^my_l^n|^d).\end{align*} $$

But the estimate

$$ \begin{align*}|x_j^my_j^n|^d\ge \frac 12 \frac{|x^my^n|^d}{1+j|x^my^n|^d}\end{align*} $$

from assertion 3 of Proposition 2.2 shows that $\sum |x_l^my_l^n|^d=+\infty $ , so $|y_j|\to \infty $ , which is a contradiction. In the same way, up to reducing $\delta '$ , we can prove that the negative orbit of any $(x,y)\in \mathcal {U}\backslash \{xy=0\}$ leaves $\mathcal {U}$ .

Suppose now that F is of the form (A). Again as in §3, up to a linear change of coordinates, we have

$$ \begin{align*} F(x,y)=(x+x^{M+1}[-1/M+O(x,y)], y+x^M[by+O(x,y^2)]), \end{align*} $$

where $\operatorname {\mathrm {Re}} b>0$ . As explained in the previous section, for some $\varepsilon _0>0$ there exists a holomorphic map $u_k:S_k(M,\varepsilon _0,\pi /4)\to \mathbb C$ such that in the sectorial coordinates $(x,z^k)=(x,y-u_k(x))$ defined in $S_k(M,\varepsilon _0,\pi /4)\times \mathbb {C}$ we can write

$$ \begin{align*}F(x,z^k)=(x+x^{M+1}[-1/M+O_1(x,z^k)], z^k+x^Mz^k[b+O_1(x,z^k)]).\end{align*} $$

Since $z^k_1=z^k(1+[b+O_1(x,z^k)]x^M)$ and $\operatorname {\mathrm {Re}} b>0$ , as in the previous case we can find constants $\varepsilon , \delta ,\nu>0$ and $\theta \in (0,\pi /4)$ such that, for each $k\in \{0,\ldots ,M-1\}$ , if $x\in S_k(M,\varepsilon ,\theta )$ with $|x|<\delta $ and $|z^k|<\delta $ then

(19) $$ \begin{align} |z^k_1|\ge |z^k|(1+\nu |x|^M). \end{align} $$

Take $\delta _0\le \delta /2$ such that $|u_k(x)|<\delta /2$ for all $|x|<\delta _0$ and $k\in \{0,\ldots , M-1\}$ . Notice that the equation

$$ \begin{align*} x_1 = x+x^{M+1}[-1/M+O(x,y)] \end{align*} $$

satisfies the hypothesis of Proposition 2.2 with $m=1$ , $n=0$ and $d=M$ , so there exists a constant $\tilde \delta _F(\theta )$ satisfying assertion 3. Set $\delta '=\min \{\delta _0,\tilde \delta _F(\theta )\}$ and

$$ \begin{align*}\mathcal{U}=\{(x,y)\in\mathbb C^2: |x|<\delta', |y|<\delta'\}.\end{align*} $$

Consider a point $(x,y)\in \mathcal {U}$ outside the fixed set $\{x=0\}$ and let us show that there exists $j\in \mathbb N$ such that $(x_j,y_j)\not \in \mathcal {U}$ . Assume by contradiction that $(x_j,y_j)\in \mathcal {U}$ for all j. Then, by assertion 3 of Proposition 2.2, we find $k\in \{0,\ldots ,M-1\}$ and $j_0\ge 0$ such that ${x_j\in S_k(M,\varepsilon ,\theta )}$ for all $j\ge j_0$ . Since $|x_j|<\delta '\le \delta _0\le \delta $ for all $j\ge 0$ , we have that

$$ \begin{align*}|z^k_j|\le |y_j|+|u_k(x_j)|<\delta'+\delta/2\le\delta\end{align*} $$

for all $j\ge 0$ . Thus, by iterated applications of (19) we obtain, for all $j\ge j_0$ , that

$$ \begin{align*}|z^k_j|\ge |z^k_{j_0}|\prod_{l=j_0}^{j-1}(1+\nu |x_l|^M).\end{align*} $$

But the estimate

$$ \begin{align*}|x_j|^M\ge \frac 12 \frac{|x|^M}{1+j|x|^M}\end{align*} $$

from assertion 3 of Proposition 2.2 shows that if $z^k_{j_0}\neq 0$ then $z^k_j\to +\infty $ , which is a contradiction. Hence $z^k_{j_0}=0$ , which means that the orbit of $(x,y)$ eventually lies in the parabolic curve $z^k=0$ of F in the domain $S_k(M,\varepsilon ,\pi /4)\times \mathbb {C}$ . For each ${k\in \{0,\ldots ,M-1\}}$ , let $\mathcal {P}_k$ be the set of points in $\mathcal {U}$ attracted by $F|_{\mathcal U}$ into the parabolic curve of F in $S_k(M,\varepsilon ,\pi /4)\times \mathbb {C}$ ; it is not difficult to see that $\mathcal P_k$ is a one-dimensional complex submanifold of $\mathcal U$ , and we have shown that every point in $\mathcal U$ outside the fixed set $\{x=0\}$ and outside the manifold $\mathcal P^+ =\mathcal P_0 \cup \cdots \cup \mathcal P_{M-1}$ has a finite positive orbit in $\mathcal U$ . In the same way, up to reducing $\delta '$ , if $\mathcal P_k^-$ is the one-dimensional complex submanifold of $\mathcal U$ of points attracted by $F^{-1}|_{\mathcal U}$ into the parabolic curve of $F^{-1}$ in the domain $S_k^-(M,\varepsilon , \pi /4)\times \mathbb {C}$ , then every point in $\mathcal U$ outside the fixed set $\{x=0\}$ and outside the manifold $\mathcal P^-=\mathcal P_0^- \cup \cdots \cup \mathcal P_{M-1}^-$ has a finite negative orbit in $\mathcal U$ . This concludes the proof of Theorem 1.3.

Acknowledgements

The first author was partially supported by Ministerio de Ciencia e Innovación, Spain, PID2019-105621GB-I00; the second author by Vicerrectorado de Investigación de la Pontificia Universidad Católica del Perú.

A Appendix. Resolution theorem for biholomorphisms

The resolution theorem for two-dimensional biholomorphisms stated in the introduction is valid not only for biholomorphisms tangent to the identity, but more generally for unipotent biholomorphisms, and is based on the corresponding theorem for vector fields and foliations in $\mathbb C^2$ . A formal vector field X in $(\mathbb {C}^2,0)$ can be written, in a unique way up to multiplication by a unit, as $X=f(A({\partial }/{\partial x})+B({\partial }/{\partial y})),$ where $f,A,B\in \mathbb C[[x,y]]$ and A and B have no common factor. The vector field $\operatorname {\mathrm {Sat}} X=A({\partial }/{\partial x})+B({\partial }/{\partial y})$ is called the saturation of X. If f is not a unit, we say that $\operatorname {\mathrm {Sing}} X=\sqrt {(f)}$ is the singular locus of X; if f is a unit, we say that X is saturated and define $\operatorname {\mathrm {Sing}} X=\{0\}$ if A and B are not units and $\operatorname {\mathrm {Sing}} X=\emptyset $ otherwise. An irreducible formal curve $(g)$ in $(\mathbb {C}^2,0)$ is said to be a separatrix of a formal vector field X if $X(g)\in (g)$ . If X is not singular, its formal integral curve through the origin is its only separatrix. The branches of the singular locus of X are separatrices, which are called fixed.

We say that a saturated singular vector field X in $(\mathbb {C}^2,0)$ is reduced if the eigenvalues $\unicode{x3bb} _{1},\unicode{x3bb} _{2}$ of its linear part satisfy $\unicode{x3bb} _{1}\neq 0$ and $\unicode{x3bb} _{2}/\unicode{x3bb} _{1}\not \in \mathbb Q_{>0}$ ; if $\unicode{x3bb} _2\neq 0$ we say that X is non-degenerate, otherwise X is called a saddle-node. A reduced vector field X has exactly two formal separatrices, which are non-singular and transverse, and each one is tangent to an eigenspace of the linear part of X. The resolution theorem for vector fields (see [Reference Camacho, Lins Neto and Sad4, Reference Seidenberg14]) asserts the following (throughout this section, if $\mathfrak g$ is any analytic or formal object, we denote by $\mathfrak g_p$ its germ at the point p).

Theorem A.1. Let X be a singular formal vector field in $(\mathbb {C}^2,0)$ . There exist a finite composition of blow-ups $\pi \colon (M,E)\to (\mathbb {C}^2,0)$ , a formal vector field $\tilde {X}$ along E with $\pi _*\tilde {X}=X$ , and finitely many points $p_1,\ldots , p_k\in E$ such that $\operatorname {\mathrm {Sat}}\tilde X_{p_1},\ldots ,\operatorname {\mathrm {Sat}}\tilde X_{p_k}$ are reduced and $\operatorname {\mathrm {Sat}}\tilde X_p$ is not singular for any $p\in E\backslash \{p_1,\ldots ,p_k\}$ .

The set E, called the exceptional divisor, is a finite union of smooth rational curves with normal crossings; we say that a point in E is a corner if it is the intersection of two components of D. Up to composing $\pi $ with some additional blow-ups, we can assume that the family of separatrices of X, even if it is infinite, is desingularized by $\pi $ :

  1. (a) the strict transform of each separatrix is a non-singular curve at a non-corner point of E and is transverse to E;

  2. (b) the strict transforms of different separatrices are curves at different points of E.

Any map $\pi $ as above is called a resolution of X and we also say that $\tilde X$ is a resolution of X. Any further blow-up at a singular point of $\tilde X$ in E gives another resolution of X. As may be expected, there exists a unique minimal resolution of X in the sense that any other resolution is obtained from the minimal one by performing finitely many additional blow-ups. If X is a nilpotent vector field in $(\mathbb {C}^2,0)$ (that is, its linear part is nilpotent) and $\tilde X$ is a resolution of X, then $\tilde X_p$ is nilpotent for any point $p\in E$ . In particular, if X is nilpotent then the singular points of $\tilde X$ in E are not isolated: an isolated singularity p of $\tilde X$ would be saturated and reduced, so $\tilde X_p$ would not be nilpotent. Therefore the set $\operatorname {\mathrm {Sing}}_E\tilde X$ of singular points of $\tilde X$ in E is a union of components of E. We define the singular locus $\operatorname {\mathrm {Sing}}\tilde X$ of $\tilde X$ as the union of $\operatorname {\mathrm {Sing}}_E\tilde X$ with the strict transform of $\operatorname {\mathrm {Sing}} X$ by $\pi $ . It is easy to see that $(\operatorname {\mathrm {Sing}}\tilde X)_p=\operatorname {\mathrm {Sing}} \tilde X_p$ for all $p\in E$ . Observe that $E\cup \operatorname {\mathrm {Sing}}\tilde X$ is the transform of $\operatorname {\mathrm {Sing}} X$ , so it is a finite union of smooth curves with normal crossings, and so is $\operatorname {\mathrm {Sing}}\tilde X$ .

Lemma A.2. Let $\tilde X$ be a resolution of X and let $p\in E$ such that $\operatorname {\mathrm {Sat}}\tilde {X}_p$ is singular (hence a reduced singularity). Then each branch of $(E\cup \operatorname {\mathrm {Sing}} \tilde X)_p$ is one of the two separatrices of $\operatorname {\mathrm {Sat}}\tilde {X}_p$ .

Proof. Let $S_1$ and $S_2$ be the separatrices of $\operatorname {\mathrm {Sat}}\tilde {X}_p$ . If both $S_1$ and $S_2$ were not contained in E, they would be the strict transforms of different separatrices of X passing through the same point in E, contradicting (a), so we can assume that $S_1$ is contained in a component of E. Suppose that there is a branch S of $(E\cup \operatorname {\mathrm {Sing}} \tilde X)_p$ different from $S_1$ and $S_2$ . If S is a component of E then p is a corner and $S_2$ is the strict transform of a separatrix of X passing through p, which contradicts (a). If S is a branch of $\operatorname {\mathrm {Sing}}\tilde X_p$ not contained in E then S is the strict transform of a separatrix of X, so by (a) p is not a corner and then $S_2$ is also the strict transform of a separatrix of X passing through p, contradicting (b).

If $\tilde X$ is a resolution of X, we classify the components of the exceptional divisor into two types.

  1. (1) A component D of E is invariant if for some point $p\in D$ the germ $D_p$ is a separatrix of $\operatorname {\mathrm {Sat}}\tilde X_p$ . In this case the same happens for any other point in D.

  2. (2) If a component D of E is not invariant, we say that it is dicritical. In this case ${D\subset \operatorname {\mathrm {Sing}}\tilde X}$ and, as we will see next, the vector field $\operatorname {\mathrm {Sat}}\tilde X_p$ is non-singular and transverse to D for all $p\in D$ , and any other component of E intersecting D is invariant.

Let us show the assertions in (2). By Lemma A.2, if $\operatorname {\mathrm {Sat}}\tilde X_p$ were singular for some $p\in D$ then D would be invariant, so $\operatorname {\mathrm {Sat}}\tilde X_p$ is not singular and its formal integral curve C is different from $D_p$ because D is not invariant. If p is not a corner, then C is the strict transform of a separatrix of X, so C is transverse to D; if p is a corner, from (a) we conclude that $C=D^{\prime }_p$ , where $D'$ is the other component of E through p and therefore $\operatorname {\mathrm {Sat}}\tilde X_p$ is transverse to D. This also shows that any component $D'$ of E intersecting D is invariant.

Consider now a unipotent biholomorphism F, that is, $DF(0)=I+N$ where I is the identity and N is nilpotent. In a formal sense, F is the time-one flow of a unique formal vector field in $(\mathbb {C}^2,0)$ , denoted $\log F$ , which is singular at the origin and has N as linear part. In particular, if F is tangent to the identity then $\log F$ has order at least two. Moreover, the fixed point set of F coincides with the singular locus of $\log F$ , which is therefore convergent. If $\pi $ is the blow-up at the origin, the map $\tilde F=\pi ^{-1}\circ F\circ \pi $ is a biholomorphism in a neighborhood of $E=\pi ^{-1}(0)$ which leaves E invariant, and satisfies $\log \tilde F_p=\tilde X_p$ for any fixed point $p\in E$ of $\tilde F$ .

Theorem A.3. Let F be a unipotent biholomorphism in $(\mathbb {C}^2,0)$ , let $\pi $ be a resolution of $\log F$ and let $\tilde F=\pi ^{-1}\circ F\circ \pi $ be the transform of F by $\pi $ . Then, if $p\in E=\pi ^{-1}(0)$ is a fixed point of $\tilde F$ , the germ $\tilde {F}_p$ is reduced, according to Definition 1.1. If F is tangent to the identity and $\pi $ is the minimal resolution of $\log F$ , then E is fixed pointwise by $\tilde F$ .

Proof. Let $\tilde X$ be the transform of $X=\log F$ by $\pi $ and let $p\in E$ be a fixed point of $\tilde F$ , so $p\in \operatorname {\mathrm {Sing}}\tilde X$ .

Suppose first that p belongs to a dicritical component D of E. Then $D\subset \operatorname {\mathrm {Sing}}\tilde X_p$ and $\operatorname {\mathrm {Sat}} \tilde X_p$ is non-singular and transverse to D. Since $(D\cup \operatorname {\mathrm {Sing}} \tilde X)_p$ is smooth or has two smooth transverse branches, we can take holomorphic coordinates $(x,y)$ at p such that $D=\{x=0\}$ and such that $\{y=0\}$ is the other branch of $\operatorname {\mathrm {Sing}} \tilde X_p$ if it exists, so $\operatorname {\mathrm {Sing}}\tilde X_p\subset \{xy=0\}$ . Then, up to rescaling the coordinates, we have

$$ \begin{align*}\tilde X_p=x^My^N\bigg[(1+\tilde A(x,y))\dfrac{\partial}{\partial x}+ \tilde B(x,y)\dfrac{\partial}{\partial y}\bigg],\end{align*} $$

where $\operatorname {\mathrm {ord}}\tilde A\ge 1$ , $M\ge 1$ , $N\ge 0$ and $(M,N)\neq (1,0)$ because $\tilde X_p$ is nilpotent. Therefore its time-one flow $\tilde F_p$ will be of the form (i) once we show that $\tilde B\in (y)$ if $N\ge 1$ . Suppose that $N\ge 1$ , so $\{y=0\}\subset \operatorname {\mathrm {Sing}}\tilde X_p$ , and let C be the formal integral curve of $(1+\tilde A)({\partial }/{\partial x})+\tilde B({\partial }/{\partial y})$ through the origin. If $\{y=0\}$ is a component of E then p is a corner and, in view of (a), necessarily $C=\{y=0\}$ and therefore $\tilde B\in (y)$ . If $\{y=0\}$ is not a component of E, then it is the strict transform of a (fixed) separatrix of X so, in view of (b), it has to coincide with C and again $\tilde B\in (y)$ .

Suppose now that p does not belong to a dicritical component of E. We assume first that $\operatorname {\mathrm {Sat}}\tilde X_p$ is not singular. Take a component D of E such that $p\in D$ . As in the previous case, we have holomorphic coordinates $(x,y)$ at p such that $D=\{y=0\}$ and ${\operatorname {\mathrm {Sing}}\tilde X_p\subset \{xy=0\}}$ . Since D is invariant, $\{y=0\}$ is the formal integral curve of $\operatorname {\mathrm {Sat}}\tilde X_p$ through p, so we can write $\operatorname {\mathrm {Sat}}\tilde X_p=(1+\hat A)({\partial }/{\partial x})+\hat B({\partial }/{\partial y}) $ with $\hat B\in (y)$ . Then, up to rescaling the coordinates, we have

$$ \begin{align*}\tilde X_p=x^My^N\bigg[(1+\tilde A)\dfrac{\partial}{\partial x}+\tilde B\dfrac{\partial}{\partial y}\bigg],\end{align*} $$

where $M,N\ge 0$ and $M,N\notin \{(0,0),(1,0)\}$ because $\tilde X_p$ is singular and nilpotent, so $\tilde F_p$ is of the form (i).

Assume now that $\operatorname {\mathrm {Sat}}\tilde X_p$ is singular, hence reduced. We suppose first that it is a saddle-node. Let S be the separatrix of $\operatorname {\mathrm {Sat}}\tilde X_p$ that is tangent to the eigenspace associated to the non-zero eigenvalue of the linear part of $\operatorname {\mathrm {Sat}}\tilde X_p$ , and let $S_0$ be the other separatrix. Let $(x,y)$ be holomorphic coordinates at p such that $\{x=0\}$ and $\{y=0\}$ are respectively tangent to $S_0$ and S, so we can write $\operatorname {\mathrm {Sat}}\tilde X_p=(x+\hat A)({\partial }/{\partial x})+\hat B({\partial }/{\partial y})$ , where $\operatorname {\mathrm {ord}} \hat A, \operatorname {\mathrm {ord}} \hat B\ge 2$ and $x+\hat A$ and $\hat B$ have no common factors. We know that $\operatorname {\mathrm {Sing}}\tilde X_p$ has one or two branches, which by Lemma A.2 are contained in $\{S,S_0\}$ . Thus, we can assume that the coordinates are chosen in such a way that if $S_0\subset \operatorname {\mathrm {Sing}}\tilde X_p$ then $S_0=\{x=0\}$ , so $\hat A\in (x)$ , and if $S\subset \operatorname {\mathrm {Sing}}\tilde X_p$ then $S=\{y=0\}$ , so $\hat B\in (y)$ ; hence $\operatorname {\mathrm {Sing}}\tilde X_p\subset \{xy=0\}$ . Thus, up to rescaling the coordinates, we have

$$ \begin{align*}\tilde X_p=x^My^N\bigg[(x+\tilde A)\dfrac{\partial}{\partial x}+\tilde B\dfrac{\partial}{\partial y}\bigg],\end{align*} $$

where $M+N\ge 1$ , $\operatorname {\mathrm {ord}} \tilde A, \operatorname {\mathrm {ord}}\tilde B\ge 2$ , $\tilde A\in (x)$ if $M\ge 1$ , $\tilde B\in (y)$ if $N\ge 1$ and $x+\tilde A$ and $\tilde B$ have no common factors. Hence, $\tilde F_p$ is of the form (iii).

Finally, suppose that $\operatorname {\mathrm {Sat}}\tilde X_p$ is non-degenerate. Since $\operatorname {\mathrm {Sing}} \tilde X_p$ contains at least one branch, in suitable coordinates $(x,y)$ we have $\{x=0\}\subset \operatorname {\mathrm {Sing}}\tilde X_p$ so, by Lemma A.2, the curve $\{x=0\}$ is a separatrix of $\operatorname {\mathrm {Sat}}\tilde X_p$ . As in the previous case, we can assume that the other separatrix S of $\operatorname {\mathrm {Sat}}\tilde X_p$ is tangent to $\{y=0\}$ and that $S=\{y=0\}\subset \operatorname {\mathrm {Sing}}\tilde X_p$ if $\operatorname {\mathrm {Sing}}\tilde X_p$ has two branches, so $\operatorname {\mathrm {Sing}}\tilde X_p\subset \{xy=0\}$ . Therefore we have

$$ \begin{align*}\tilde X_p=x^My^N\bigg[(ax+\tilde Ax)\dfrac{\partial}{\partial x}+(by+\tilde B)\dfrac{\partial}{\partial y}\bigg],\end{align*} $$

where $M\ge 1$ , $N\ge 0$ , $a,b\in \mathbb {C}^*$ , $a/b\notin \mathbb {Q}_{>0}$ , $\operatorname {\mathrm {ord}}\tilde A\ge 1$ , $\operatorname {\mathrm {ord}}\tilde B\ge 2$ and $\tilde B\in (y)$ if ${N\ge 1}$ , so $\tilde F_p$ is of the form (ii).

In order to prove the last assertion of the theorem it suffices to show that if a vector field X has order at least two and $\pi $ is its minimal resolution, then $\pi ^{-1}(0)\subset \operatorname {\mathrm {Sing}}\tilde X$ . Suppose that this property holds when the minimal resolution of X is achieved with fewer than ${n\in \mathbb {N}}$ blow-ups, and let X be a formal vector field with $\operatorname {\mathrm {ord}} X\ge 2$ whose minimal resolution is obtained with n blow-ups. Let $\sigma $ be the blow-up at the origin and let $\hat X$ be the transform of X by $\sigma $ . Since $\operatorname {\mathrm {ord}} X\ge 2$ , we have that $\hat X$ vanishes on $D=\sigma ^{-1}(0)$ with order $\nu \ge 1$ if D is invariant or $\nu \ge 2$ if D is dicritical. So D will be in the singular locus of the resolution of X and, in view of the inductive hypothesis, it is enough to show that $\hat X$ has order at least two at each point in D that is blown up in the resolution. Let $p\in D$ be one such point. Since $\operatorname {\mathrm {ord}}\hat X_p\ge \nu $ , it suffices to consider the case where D is invariant and ${\nu =1}$ . We can also assume that $\operatorname {\mathrm {Sing}} \hat X_p =D_p$ and that $\hat X_p$ vanishes on $D_p$ with multiplicity one, because otherwise $\operatorname {\mathrm {ord}}\hat X_p\ge 2$ . Then, if $(x,y)$ are holomorphic coordinates at p such that $D_p=\{y=0\}$ , we have

$$ \begin{align*}\hat X_p=y\bigg[(a+\tilde A)\dfrac{\partial}{\partial x}+(b+\tilde B)\dfrac{\partial}{\partial y}\bigg],\end{align*} $$

where $\operatorname {\mathrm {ord}} \tilde A,\operatorname {\mathrm {ord}} \tilde B\ge 1$ , and necessarily $b=0$ because $\hat X_p$ is nilpotent. If $a\neq 0$ , we see that $\hat X_p$ is actually in final form, so no further blow-up at p would be necessary. Therefore $a=0$ and $\operatorname {\mathrm {ord}}\hat X_p\ge 2$ .

Remark A.4. The reduced models (i), (ii) and (iii) of Definition 1.1 correspond to the standard final models for vector fields, and are not exactly the same ones that appear in the resolution theorem in [Reference Abate1]. Our set of reduced fixed points is stable under blow-ups, in the sense that any further blow-up will produce only reduced fixed points (for example, the dicritical fixed points considered as final models in [Reference Abate1] can be reduced by additional blow-ups to non-dicritical models). In our final models, the blow-up of a fixed point of the form (i) produces fixed points of the form (i), the blow-up of a fixed point of the form (ii) produces fixed points of the form (i) and (ii), and the blow-up of a fixed point of the form (iii) produces fixed points of the form (i), (ii) and (iii).

References

Abate, M.. The residual index and the dynamics of holomorphic maps tangent to the identity. Duke Math. J. 107(1) (2001), 173207.CrossRefGoogle Scholar
Arizzi, M. and Raissy, J.. On Écalle–Hakim’s theorems in holomorphic dynamics. Frontiers in Complex Dynamics (Princeton Mathematical Series, 51). Ed. A. Bonifant, M. Lyubich and S. Sutherland. Princeton University Press, Princeton, NJ, 2014, pp. 387449.CrossRefGoogle Scholar
Brochero Martínez, F. E.. Groups of germs of analytic diffeomorphisms in $\left({\mathbb{C}}^2,0\right)$ . J. Dyn. Control Syst. 9 (2003), 132.CrossRefGoogle Scholar
Camacho, C., Lins Neto, A. and Sad, P.. Topological invariants and equidesingularization for holomorphic vector fields. J. Differential Geom. 20 (1984), 143174.CrossRefGoogle Scholar
Écalle, J.. Les fonctions résurgentes. Tome III . L’équation du pont et la classification analytique des objects locaux (Publications Mathématiques d’Orsay, 85-5). Université de Paris-Sud, Département de Mathématiques, Orsay, 1985.Google Scholar
Fatou, P.. Sur les équations fonctionelles. Bull. Soc. Math. France 47 (1919), 161271.CrossRefGoogle Scholar
Hakim, M.. Transformations tangent to the identity. Stable pieces of manifolds. Prépublication d’Orsay, no. 30, 1997.Google Scholar
Hakim, M.. Analytic transformations of $\left({\mathbb{C}}^{\mathrm{p}},0\right)$ tangent to the identity. Duke Math. J. 92(2) (1998), 403428.CrossRefGoogle Scholar
Leau, L.. Étude sur les équations fonctionelles à une ou à plusieurs variables. Ann. Fac. Sci. Toulouse Sci. Math. Sci. Phys. 11 (1897), 1110.Google Scholar
Lisboa, L.. Sobre biholomorfismos com órbitas finitas e conjuntos analíticos invariantes. PhD Thesis, Universidade Federal Fluminense, 2019.Google Scholar
López-Hernanz, L.. Summable formal invariant curves of diffeomorphisms. Ergod. Th. & Dynam. Sys. 32(1) (2012), 211221.CrossRefGoogle Scholar
López-Hernanz, L. and Rosas, R.. Characteristic directions of two-dimensional biholomorphisms. Compos. Math. 156(5) (2020), 869880.CrossRefGoogle Scholar
Milnor, J.. Dynamics in One Complex Variable (Annals of Mathematics Studies, 160). Princeton University Press, Princeton, NJ, 2006.Google Scholar
Seidenberg, A.. Reduction of the singularities of the differential equation $Ady=Bdx$ . Amer. J. Math. 90 (1968), 248269.CrossRefGoogle Scholar
Vivas, L.. Degenerate characteristic directions for maps tangent to the identity. Indiana Univ. Math. J. 61(6) (2012), 20192040.CrossRefGoogle Scholar